Vous êtes sur la page 1sur 133

MARKOV CHAIN STOCHASTIC SEISMIC INVERSION FOR

ROCK PROPERTIES ESTIMATION

--------------------------------------------------------

A Dissertation Presented to

The Faculty of the Department Earth and Atmospheric Sciences

University of Houston

--------------------------------------------------------

In Partial Fulfillment

of the Requirements for the Degree

Doctor of Philosophy

-------------------------------------------------------

By

Carlos Cobos

December 2014
MARKOV CHAIN STOCHASTIC SEISMIC INVERSION FOR

ROCK PROPERTIES ESTIMATION

____________________________________

Carlos Cobos

APPROVED:

____________________________________

Dr. John Castagna, Chairman

____________________________________

Dr. Evgeny M. Chesnokov, Member

____________________________________

Dr. Aibing Li, Member

____________________________________

Dr. Franklin Ruiz, Member

___________________________________

Dean, College of Natural Sciences and


Mathematics

ii
ACKNOWLEDGMENTS

I would like to express my gratitude to John Castagna for being an

outstanding advisor, who has provided assistance, guidance and support. John is

much more than an academic advisor; he is truly a “maestro”. I would also like to

thank Aibing Li, Evgeny Chesnokov and Franklin Ruiz, for their readiness to

serve on my oral exam committee and for their thorough review and constructive

comments on my dissertation and for their tremendous help and support to finish

it. Special thanks go to Aibing Li for her help and guidance as graduate student

advisor in Geophysics.

I want to thank Sylvia Marshal for her patience and help in all these great

years. I am also very grateful to Anne Decour and Hua-wei Zhou; without their

support I would not have been able to complete this dissertation.

I have enjoyed and learned from many discussions with Igor Escobar,

Ezequiel Gonzalez, Andres Mantilla, Per Avseth, Mario Gutierrez, Isao

Takahashi, Manuel Gonzalez and Marcelo Benabentos. I am also very grateful to

Tapan Mukerji and De-Hua Han; I cannot thank them enough for all the support

they have given me all these years.

I have been fortunate to share my personal time with a number of great

friends. Among them the most memorable are Carlos Moreno, Javier Villegas,

Tony De Lilla, Maria Villegas, Alexandra Marquez, Carlos Lopez, Karen Romero,

iii
Laszlo Benkovics, Ana Paredes, Freddy Obregon Roberto Varade and Noemi

Ordonez.

I have infinite gratitude to my parents Gracia and Carlos, my grandfather

Carlos Eduardo, my siblings, Gracia, Carlos Jose, Ligia and Francisco Jesus,

and the rest of my family for their continued trust and support.

Finally, I dedicate this dissertation to my wife Carmen and my children

Alejandra, Carla and Carlos Eduardo. They have been the energy and motivation

that made possible the culmination of my Ph.D.

iv
MARKOV CHAIN STOCHASTIC SEISMIC INVERSION FOR

ROCK PROPERTIES ESTIMATION

---------------------------------------------------

An Abstract Dissertation

Presented to

The Faculty of the Department Earth and Atmospheric Sciences

University of Houston

--------------------------------------------------------

In Partial Fulfillment

of the Requirements for the Degree

Doctor of Philosophy

-------------------------------------------------------

By

Carlos Cobos

December 2014

v
ABSTRACT

Our purpose is to introduce a stochastic seismic inversion algorithm based

on the Markov Chain Monte Carlo Simulation. Conventional inversion algorithms

generate elastic properties instead of the key rock properties needed to reliably

characterize the hydrocarbon potential of a given subsurface interval. In contrast,

the inversion scheme presented here generates a set of possible combinations of

rock properties that can explain seismic amplitude responses in terms of

lithology, pore structure and fluid variations. The result of the probabilistic seismic

inversion is a seismic lithofacies catalog that can describe the elastic response of

the studied subsurface interval. The main advantage of this technique is that the

results consist of many equally probable rock property models as an alternative

to multiple elastic property scenarios. Therefore, no post facto elastic-to-rock-

properties conversion is needed. The method might be used either in exploratory

areas or hydrocarbon field development. In exploratory areas, the stochastic rock

physics inversion can support the evaluation for hydrocarbon potential

considering the effects of reservoir properties on seismic signatures for different

geologic scenarios and physical conditions, with the prime goal of minimizing

uncertainties and risk. In field development areas, stochastic seismic inversion

produces multiple equally probable rock property models that can explain the real

3D seismic response; it can be used to constrain possible reservoir models used

for hydrocarbon reserve estimation and reservoir production simulation. The

probabilistic inversion algorithm was successfully tested on a set of real seismic

vi
and well-log data to demonstrate the feasibility of the estimation of critical rock

properties for hydrocarbon exploration, such as total porosity and reservoir

fraction. The real-data test results confirmed the capability of the proposed

inversion technique to accurately predict the rock properties of reservoir seismic

lithofacies, even for seismically thin layers.

vii
CONTENTS

Acknowledgments ...................................................................................... iii


Abstract......................................................................................................vi
List of Tables ............................................................................................. x
List of Figures ............................................................................................xi
1. Introduction ....................................................................................... 1
1.1. Motivation ................................................................................... 1
1.2. Chapter Layout ........................................................................... 4
2. Markov Chain Monte Carlo Simulation.............................................. 7
2.1. Introduction................................................................................. 7
2.2. Random Variables ...................................................................... 7
2.3. Markov Process and Markov Chains ........................................ 12
2.4. Markov Chain Monte Carlo Methods ........................................ 16
2.5. Metropolis–Hastings Algorithm ................................................. 18
3. The Statistical Rock Physics Model ................................................ 22
3.1. Introduction............................................................................... 22
3.2. Rock Physics Models ............................................................... 22
3.3. Deterministic Versus Statistical Models .................................... 24
3.4. Probabilistic Rock Physics Models ........................................... 25
3.5. The multidimensional Non-Parametric Model ........................... 27
3.6. Hybrid Rock Physics Model Example ....................................... 29
3.7. Using MCMCS to Explore Model Space ................................... 36
4. 1D Markov chain stochastic seismic inversion ................................ 38
4.1. Introduction............................................................................... 38
4.2. Proposed Inversion Algorithm .................................................. 39
4.3. Possible Scenario Generation .................................................. 41
4.4. Synthetic Seismic Generation .................................................. 44
4.5. The Misfit Function ................................................................... 48
5. Synthetic Examples ............................................................................. 51
5.1. Introduction ................................................................................... 51

viii
5.2. Synthetic Test 1: Cemented Sand ................................................. 52
5.3. Synthetic Test 2: High Porosity Sands .......................................... 66
5.4. Conclusions................................................................................... 77
6. 1D Real Data Application ................................................................ 79
6.1. Introduction............................................................................... 79
6.2. Seismic Data Conditioning ....................................................... 81
6.3. The Model Space Generation................................................... 83
6.4. The Seismic Stochastic Inversion ............................................. 85
6.5. Analysis .................................................................................... 90
6.6. Conclusions .............................................................................. 95
7. 3D Real Data Application ................................................................ 96
7.1. Introduction............................................................................... 96
7.2. The Kriging Estimator ............................................................... 96
7.3. 3D Real Data Application ......................................................... 99
7.4. Convergence Criteria ............................................................. 109
7.5. Algorithm Efficiency ................................................................ 111
6.7. Conclusions ............................................................................ 112
References ............................................................................................ 113

ix
LIST OF TABLES

Table 2.1: Probability associated to the possible outcomes of random process. .. 9


Table 2.2: Cumulative probability associated to the possible outcomes of the RV
X. ........................................................................................................................ 10
Table 2.3: Markov chain for initial price variation “same”. ................................... 15
Table 2.4: Markov chain for initial price variation “decrease”. ............................. 16
Table 3.1: Vp statistical summary table for seismic lithofacies A and B. ............ 27
Table 3.2: Fluid properties generated using Batzle and Wang (1992) fluid
properties empirical relations. ............................................................................. 32
Table 4.1: Example of the Fisher and Yates shuffle algorithm. The permuted
interval ranges from 1 to 5. The red cells correspond to the number that is
selected in each iteration. ................................................................................... 43
Table 5.1: Comparison between the synthetic data and the depths obtained by
the inversion. ...................................................................................................... 75
Table 6.1: Comparison between the real and the estimated lithology top values
for probability larger than 0.2. ............................................................................. 91
Table 6.2: Comparison between the real and the estimated fluid contact depths.
........................................................................................................................... 93

x
LIST OF FIGURES

Figure 2.1: Probability density function of the discrete random variable described
in Table 2.1. .......................................................................................................... 9
Figure 2.2: Cumulative probability density function of the RV described in Table
2.2. ..................................................................................................................... 11
Figure 2.3: Schematic representation of the transition probability matrix. .......... 15
Figure 2.4: Markov Chain evolution after 500, 1,000, 5,000, 10,000, 25,000 and
50,000 steps. The equilibrium distribution is a Gaussian with zero mean and
variance of one. .................................................................................................. 18
Figure 2.5: Two-dimensional RW with 10,000 points.......................................... 19
Figure 3.1: Deterministic (left) versus stochastic models (right). The stochastic
model generates a set of possible values instead of a simple outcome. ............ 25
Figure 3.2: P-wave velocity histograms for two different lithofacies.................... 26
Figure 3.3: Three-dimensional projections of a ten-dimensional Gaussian PDF
sampled with 1,000 (upper left), 10,000 (upper right), 100,000 (lower left) and
1,000,000 (lower right). ....................................................................................... 29
Figure 3.4: 2D Projections of the multivariate probability distribution function. A)
Bivariate P-wave velocity versus S-wave velocity PDF; B) Bivariate P-wave
velocity versus bulk density PDF; C) Bivariate S-wave velocity versus bulk
density PDF; and D) Bivariate Porosity versus bulk density PDF. ...................... 31
Figure 3.5: 2D Projections of the modeled multivariate probability distribution
function. A) Bivariate P-wave velocity versus Vs PDF; B) Bivariate P-wave
velocity versus bulk density PDF; C) Bivariate S-wave velocity versus bulk
density PDF; and D) Bivariate Porosity versus bulk density PDF. ...................... 33
Figure 3.6: Elastic properties comparison between original saturation (left) and
gas replaced saturation (right). ........................................................................... 34
Figure 3.7: Example of theoretical rock physics models application to measured
P-wave velocity samples. The fluid replacement, cement modeling and porosity
perturbation effects are shown by red, green and black arrows. ........................ 35
Figure 3.8: Example of theoretical rock physics models application to measured
P-wave velocity samples. The probability distribution functions for the fluid
replacement, cement modeling and porosity perturbation effects are in red, green
and black, respectively. ...................................................................................... 35
Figure 3.9: Comparison of an original bivariate distribution (left) and the MCMCS
distribution (right). ............................................................................................... 37

xi
Figure 3.10: Mean square error plot for the MCMCS of the example shown in
Figure 3.9. .......................................................................................................... 37
Figure 4.1: The inverse and forward problems are complementary mathematical
problems. ............................................................................................................ 38
Figure 4.2: Simplified 1D stochastic inversion workflow. .................................... 41
Figure 4.3: Schematic representation of the possible scenario generation. ....... 43
Figure 4.4: Comparison among the depth, time and regularized time 1-D P-wave
velocity arrays. The black circles show how the regularization in the time domain
can affect the elastic properties of the modeled subsurface. .............................. 46
Figure 4.5: Schematic representation of the synthetic traces generation
algorithm. ............................................................................................................ 48
Figure 4.6: Misfit calculation example. In this synthetic example the cumulative
L2-norm shows the intervals with larger misfit. .................................................... 49
Figure 5.1: Model 1 synthetic well logs showing (from left to right) facies
indicator, shale volume, P-wave velocity, bulk density, S-wave velocity, porosity,
P-wave impedance and S-wave impedance. ...................................................... 52
Figure 5.2: Synthetic model 1 pore fluids velocity and density profiles. .............. 54
Figure 5.3: Final seismic lithofacies multivariate PDF projections for the Vp-Vs
(Upper), Vp-Rhob (Center), and PhiT-Vp (Lower) spaces. Green, blue, brown
and red contours denote shales, brine sands, oil sands and gas sands,
correspondingly. Some empirical models are plotted for comparison including
Castagna’s mudrock line, Castagna’s 1993, Han’s 1986, Gardner’s sand,
Gardner’s shale, Voigt bound, Reuss bound and Nu’s modified Voigt bound for
quartz mixture. .................................................................................................... 55
Figure 5.4: Model 1 angle dependent reflectivity profiles.................................... 56
Figure 5.5: Ricker wavelet with a central frequency of 25 Hz. ............................ 57
Figure 5.6: Synthetic angle gather generated from the synthetic model 1. ......... 57
Figure 5.7: 2D Projections of the multivariate probability distribution function for
sands (left) and shales (right) calculated from the synthetic model 1. ................ 58
Figure 5.8: Real seismic trace (left) and possible models traces (right)
comparison. The selected traces have L2-norms smaller than 0.9. .................... 59
Figure 5.9: Synthetic versus modeled waveform comparison for four possible
models with α<0.9. ............................................................................................. 60
Figure 5.10: Probable models with α<0.9. The black, red, yellow and green lines
represent the top of upper sand, base of upper sand, top of lower sand and base
of lower sand, respectively. ................................................................................ 61
Figure 5.11: Sandstone fraction for probable models with α< 0.9....................... 62

xii
Figure 5.12: Total porosity of the probable models with α<0.9. The black, red,
yellow and green lines represent the top of upper sand, base of upper sand, top
of lower sand and base of lower sand, respectively. .......................................... 63
Figure 5.13: Volume of shale probable models with α<0.9. The black, red, yellow
and green lines represent the top of upper sand, base of upper sand, top of lower
sand and base of lower sand, respectively. ........................................................ 63
Figure 5.14: Effective porosity probable models with α<0.9. The black, red,
yellow and green lines represent the top of upper sand, base of upper sand, top
of lower sand and base of lower sand, respectively. .......................................... 64
Figure 5.15: Elastic models with α<0.9. The black, red, yellow and green lines
represent the top of upper sand, base of upper sand, top of lower sand and base
of lower sand respectively. ................................................................................. 64
Figure 5.16: Normalized histograms comparison between the rock properties of
the synthetic model and the probable solutions, for sand’s shale volume (upper
left), sand’s total porosity (lower left), shale’s shale volume (upper right) and
shale’s total porosity (lower right). Red and blue bars correspond to synthetic
model and solutions, respectively. ...................................................................... 65
Figure 5.17: Probable models with α<0.045. The black, red, yellow and green
lines represent the top of upper sand, base of upper sand, top of lower sand and
base of lower sand, respectively. ........................................................................ 66
Figure 5.18: Model 2 synthetic well logs showing (from left to right) facie
indicator, shale volume, porosity, P-wave velocity, bulk density and S-wave
velocity. Black, blue and red lines correspond to in situ, brine and gas
saturations. ......................................................................................................... 68
Figure 5.19: Synthetic model 2 pore fluids velocity (left) and density profiles
(right). Blue and red lines correspond to brine and gas, respectively. ................ 68
Figure 5.20: Final seismic lithofacies multivariate PDF projections for the Vp-Vs
(Upper), Vp-Rhob (Center), and PhiT-Vp (Lower) spaces. Grey, blue, and red
contours denote shales, brine sands and gas sands, correspondingly. Some
empirical models are plotted for comparison, including Castagna’s Mudrock line,
Castagna’s 1993, Han’s 1986, Gardner’s sand, Gardner’s shale, Voigt bound,
Reuss bound and Nu’s modified Voigt bound for quartz mixture. ....................... 69
Figure 5.21: Model 2 angle-dependent reflectivity profiles. ................................ 70
Figure 5.22: Synthetic angle gather generated from synthetic model 2. ............. 71
Figure 5.23: 2D Projections of the multivariate probability distribution function for
gas sands (left), shales (center) and brine sands (right), calculated from synthetic
model 2. .............................................................................................................. 71
Figure 5.24: Real seismic trace (left) and possible models traces (right)
comparison. The selected traces have L2-norms smaller than 0.88. .................. 72

xiii
Figure 5.25: Synthetic versus modeled waveform comparison for four possible
models with α<0.88. ........................................................................................... 73
Figure 5.26: Probable models with α<0.88. The green, black and white represent
the shales, gas sands and brine sandstone seismic lithofacies, respectively. .... 74
Figure 5.27: Depth of top of the sand (left), gas-water contact (center) and base
of the sand (right) histograms. ............................................................................ 74
Figure 5.28: Total porosity of the probable models with α<0.88. ........................ 75
Figure 5.29: Volume of shale probable models with α<0.88. .............................. 76
Figure 5.30: Effective porosity probable models with α<0.88 calculated from the
total porosity and shale volume properties. ........................................................ 76
Figure 5.31: Elastic models with α<0.88. ............................................................ 77
Figure 6.1: Well 1 logs for in situ saturation. ....................................................... 80
Figure 6.2: Near angle stack; the top of the upper and lower sands are located at
72 ms and 104 ms, respectively. ........................................................................ 80
Figure 6.3: Angle gather at the Well-1 location (Upper). The amplitude extraction
at the top of the upper sand shows a strong AVO effect (Lower). ...................... 81
Figure 6.4: Well logs converted to time showing the well to seismic tie (right
track); the black line corresponds to the measured seismic trace and blue the
synthetic seismogram. ........................................................................................ 82
Figure 6.5: Angle-dependent extracted wavelets................................................ 83
Figure 6.6: Well 1 logs after fluid substitution. Black, red and blue lines
correspond to in situ, 90% gas and 100% brine-saturated sands. ...................... 84
Figure 6.7: 2D Projections of the multivariate probability distribution function
generated from Well 1 log data and Gassmann fluid substitution. ...................... 84
Figure 6.8: Comparison between real angle traces (upper and lower left) and the
320 solutions for 0° (upper right) and 32° (lower right) incident angles. ............. 86
Figure 6.9: Probable models with α≤0.88. Grey, black and white represent
shales, gas sands and low saturation gas sands, respectively. .......................... 87
Figure 6.10: Total porosity of the 320 probable models. ..................................... 88
Figure 6.11: Volume of shale of the 320 probable models. ................................ 88
Figure 6.12: Effective porosity of the 320 probable models. ............................... 89
Figure 6.13: Possible elastic models with α≤0.88. .............................................. 89
Figure 6.14: Probability of sand vertical plot. ...................................................... 91
Figure 6.15: Total porosity and effective porosity comparison. The black lines
correspond to the well log information and the blue lines to the inverted rock
properties............................................................................................................ 92

xiv
Figure 6.16: Most expected fluid seismic lithofacies (left) and probability of fluid
type (right). ......................................................................................................... 93
Figure 6.17: Comparison between the most expected elastic models (blue line)
against well logs (black lines). ............................................................................ 94
Figure 7.1: Simplified 3D stochastic inversion workflow. .................................. 100
Figure 7.2: Anisotropic semivariogram model with a 0.1 nugget, a sill of 0.9 and
ranges of 500 meters and 50 meters. ............................................................... 100
Figure 7.3: Variance model evolution for one of the stochastic simulation
outcomes. ......................................................................................................... 102
Figure 7.4: Three possible stochastic models. Shales are represented by facies
1, gas sands by facies 2, LSGS by facies 3 and brine sands by facies 4. ........ 102
Figure 7.5: Three possible stochastic total porosity models. ............................ 103
Figure 7.6: Three possible stochastic volume of shale models......................... 103
Figure 7.7: Three possible stochastic P-wave impedance models. .................. 104
Figure 7.8: Three possible stochastic bulk density models. .............................. 104
Figure 7.9: Three possible stochastic S-wave velocity models. ........................ 105
Figure 7.10: Most expected total porosity model generated from 400 stochastic
models. ............................................................................................................. 106
Figure 7.11: Most expected volume of shale model generated from 400
stochastic models. ............................................................................................ 107
Figure 7.12: Most expected volume of P-wave velocity generated from 400
stochastic models. ............................................................................................ 107
Figure 7.13: Most expected volume of density generated from 400 stochastic
models. ............................................................................................................. 108
Figure 7.14: Most expected volume of S-wave velocity generated from 400
stochastic models. ............................................................................................ 108
Figure 7.15: Models randomly selected from the real dataset inversion. .......... 109
Figure 7.16: Comparison between real angle trace (right) and 150,000 solutions
(left) randomly selected from the five million realizations performed for the real
example inversion............................................................................................. 110
Figure 7.17: Algorithm efficiency, the proposed algorithm increase linearly with
the number of tested models. ........................................................................... 111

xv
1. INTRODUCTION

1.1. Motivation

Conventional seismic reservoir characterization (SRC) techniques were

developed more than forty years ago for exploration plays where hydrocarbon

seismic responses were relatively easy to identify. Early reservoir

characterization workflows were usually based on direct hydrocarbon indicator

(DHI) identification techniques centered on poststack seismic amplitude analysis

and amplitude versus offset (AVO) inversion. DHIs plays are usually related to

shallow high porosity reservoirs with significantly lower acoustic impedance

compared to the surrounding rocks. The associated seismic signatures of these

hydrocarbon-filled high porosity reservoirs can be anomalous high amplitude

reflections called “bright spots”.

Nowadays, conventional seismic-reservoir-characterization techniques are

becoming obsolete, since the oil industry is moving to explore areas in which the

hydrocarbons are located in deeper and more-complex reservoirs. These new

hydrocarbon plays are characterized by low-porosity and low-permeability

reservoirs with nearly undetectable pore fluid response. This means that the

future seismic-reservoir characterization goal is to predict rock properties, such

as the porosity, lithology and rock fabric of compacted and cemented porous

rock.

1
The second-most important SRC challenge is to improve seismic vertical

resolution. Currently, seismic inversion resolution is still low in new

exploration/development challenges and the improvement of seismic derived

elastic parameters is paramount for the application of reservoir properties

prediction. Techniques such as stochastic inversion have been initially developed

in an attempt to obtain from seismic data quantitative information about

subsurface rock properties on a very detailed scale.

The main limitation of current seismic-inversion algorithms is that the

solutions are usually limited to the acoustic (P-wave impedance) or elastic (P-

wave impedance, S-wave impedance, and bulk density) properties of the studied

subsurface intervals. However, these properties are not directly useful for

characterizing hydrocarbon potential. To accurately characterize the oil potential

of a possible of subsurface reservoir, we need to know key rock properties, such

as lithology, porosity and fluid type.

The goal of this dissertation is to introduce an inversion technique based

on the Markov Chain Monte Carlo Simulation (MCMCS) that can be implemented

in a stochastic petrophysical inversion scheme. Our proposed stochastic seismic

inversion’s objective is to produce a set of probable rock property volumes that

can describe the elastic seismic response of the studied interval and its

associated uncertainties.

2
The main advantage of this petrophysical inversion technique is that the

results are multiple-probable-rock-property models instead of multiple-elastic-

property scenarios. Therefore, no post facto elastic-to-rock properties conversion

is needed. The method might be used either in exploratory areas or hydrocarbon-

field development. In exploratory areas, the proposed stochastic inversion

algorithm can support the evaluation for hydrocarbon potential considering the

effects of reservoir properties on seismic signatures for different geologic

scenarios and physical conditions, with the prime goal of minimizing uncertainties

and risk. In field development areas, the stochastic inversion produces many

equally probable rock property models that can explain the real 3D seismic

response; it can be used to constrain possible reservoir models used for

hydrocarbon reserve estimation and reservoir production simulation.

In this dissertation we selected a probabilistic solution to tackle the

problem of estimating the physical properties of subsurface rocks based on

surface seismic data. According to Tarantola (2005), most probabilistic inverse

problems can be resolved using the following equation:

( ) ( ) ( ( )) (1.1)

where ( ) is the a posteriori probability density function in the model

space, K is a normalizing constant, ( ) is the a priori probability density

function in the model space, ( ) is the data probability density function and

( ) is the forward modeling operator that relates the model space to the data

3
space. The operator ( ) is usually non-linear and its role is to predict the

values of the observable parameters that would correspond to a given

model . This theoretical prediction can be described as follows:

( ) (1.2)

The main contributions of this dissertation are: 1) the introduction of the

Metropolis–Hastings MCMCS algorithm to explore multivariate non-parametric

statistical rock physics models; and, 2) the introduction of an inversion technique

that directly produces rock property volumes instead of elastic property volumes;

hence, the inversion results can be directly introduced into the geological model

of the studied area.

1.2. Chapter Layout

Chapter 2 introduces the basis of the Markov Chain Monte Carlo

Simulation (MCMCS) and its application to the proposed stochastic inversion.

First, we introduce some basic concepts of descriptive statistics. Then we define

Markov Chains and their application to Monte Carlo simulations. Finally, it is

shown how the Metropolis–Hastings MCMCS algorithm is used to explore the

multidimensional solution space in a much more effective way than systematic

sampling algorithms.

Chapter 3 focuses on the probabilistic rock physics model used as model

space for the proposed inversion scheme. We show how probability density

4
functions are generated from the integration of the available well and geological

data, and are supported by theoretical rock physics models. Second we describe

how the MCMCS is used to populate the model space for the proposed

stochastic seismic inversion.

Chapter 4 presents our novel 1D stochastic seismic inversion algorithm

based on the Metropolis–Hastings MCMCS algorithm. The inversion workflow

and its three main steps are explained: the rock properties scenario generation,

the synthetic data generation and the comparison between the real and the

synthetic traces. We illustrate how the Metropolis–Hastings MCMCS algorithm is

used to populate the rock properties of the tested models.

Chapter 5 presents two synthetic examples to illustrate the proposed

inversion technique. We show the results of the application of the proposed

inversion scheme on two different sets of synthetic data. The synthetic data sets

were generated based on real well data, with the goal of testing the inversion in

two different geological environments. The first synthetic example represents a

geological setting where the sands have larger impedance than that of the

surrounded shales. In the second example, we tested the algorithm in a high

porosity environment where the fluid property variations have a very strong effect

on the elastic properties of the studied reservoirs.

In Chapter 6, we present the first 1D application of the proposed inversion

scheme on a real set of data that include seismic and well log information. The

5
target interval is made up of four seismic lithofacies: shales, gas-filled sands, low

saturation gas sands (LSGS) and brine sands. The objective of this inversion is

to predict total porosity, volume of shale, effective porosity and fluid type

distribution in the studied interval.

Chapter 7 presents the 3D inversion of the real data set introduced in

Chapter 6. The 3D algorithm integrates the 1D stochastic inversion procedure

with a Kriging engine to produce a lateral constraint rock property model. We

analyzed the same data set used in Chapter 6 that corresponds to tertiary marine

sediments deposited on a platform margin.

6
2. MARKOV CHAIN MONTE CARLO SIMULATION

2.1.Introduction

This chapter introduces basic concepts of statistics and the Metropolis–

Hastings Monte Carlo Markov Chain Simulation (MCMCS) algorithm. This

theoretical introduction is the foundation of the seismic inversion technique based

on stochastic simulation proposed in this dissertation. The MCMCS algorithm is

used to explore the multidimensional solution space in a much more effective

way than systematic sampling algorithms. Its application is especially appropriate

in inverse Earth science problems, in which the number of variables tends to be

very large. Before introducing the MCMCS algorithm we define some basic

statistical concepts in Section 2.2. These basic concepts are random variables,

probability density functions, cumulative density functions and the most important

descriptive statistical parameters.

2.2. Random Variables

A random variable (RV) is a mathematical variable whose possible values

are the numerical outcomes of a random process. According to Papoulis

(Papoulis 1984), any function g(x) of real variables can be a RV if g(x) has the

following properties:

a) Its domain must include all the range of RV x,

b) The set {X<=x} is an event for any real number x, and

7
c) The probability of the events {X=±∞} must be zero.

Random variables can be discrete or continuous. RVs that are continuous

can assume any value within a given range. For instance, if we randomly select

several individuals from a population and measure their height, the results can

take on any value within a given range, maybe from 2 to 7 feet. In contrast,

discrete variables can only assume certain values within a defined range. For

example, when we flip a coin and count the possible outcomes, no matter the

number of flips, the possible outcomes can only be heads or tails. Because the

number of heads or tails results from a random process flipping a coin is a

discrete RV.

Random variables are defined by their probability density function (PDF)

that describes the likelihood for a given RV to take on a given value. The

probability of an RV taking a particular range of values is defined by the integral

its PDF taking over the analyzed range.

The following example will illustrate some concepts related to the random

variables. Assume an RV X that can only take the values 1, 2, 3, 4 or 5. The

probabilities associated with the possible outcomes are shown in Table 2.1.

8
Table 2.1: Probability associated to the possible outcomes of random process.

Outcome Probability
1 0.05
2 0.25
3 0.4
4 0.25
5 0.05

The probability that X is equal to 1 or 2 is the sum of two probabilities: P(X

= 1 or X = 2) = P(X = 1) + P(X = 2) = 0.05 + 0.25 = 0.3. Correspondingly, the

probability that X is greater than 1 is equal to 1 - P(X = 1) = 1 - 0.05 = 0.95. The

set of possible outcomes and their associated probabilities shown in Table 2.1

are the PDF of the RV X (Isaaks and Srivastava, 1989); Figure 2.1 shows the

PDF plot of the RV shown in Table 2.1.

0.4

0.35

0.3

0.25
Probability

0.2

0.15

0.1

0.05

0
1 2 3 4 5
Outcome

Figure 2.1: Probability density function of the discrete random variable described
in Table 2.1.

9
Suppose the random variable X may take N different values; the PDF of

the discrete RV X is then defined as:

( ) . (2.1)

The PDF of a discrete RV must satisfy the following:

a) 0 < Pi< 1 for each i


b) ∑ .

By definition, the integral of the PDF (the sum of discrete RVs) is called

the cumulative distribution function (CDF). As mentioned, the CDF is defined as

the probability that a RV X has a value less than or equal to x. The CDF of a

discrete random variable X is related to the PDF by the following expression:

{ } ∑ (2.2)

Table 2.2 shows the CDF of the RV X shown in Table 2.1. The CDF of the

RV X is displayed in Figure 2.2.

Table 2.2: Cumulative probability associated to the possible outcomes of the RV


X.

Outcome CDF
1 0.05
2 0.3
3 0.7
4 0.95
5 1

10
1

0.9

0.8

Cummulative Probability
0.7

0.6

0.5

0.4

0.3

0.2

0.1

0
1 2 3 4 5
Outcome

Figure 2.2: Cumulative probability density function of the RV described in Table


2.2.

When the PDF of an RV is known, we can calculate many parameters that

are useful to describe the statistics of the analyzed random variable. These

parameters include expected value and standard deviation of the RV. The

expected value of an RV X is defined as the weighted average of all the

outcomes; each one is weighted by the probability of occurrence of that outcome:

{ } ∑ (2.3)

Based on the information shown in Table 2.1 we can calculate the

expected value of the RV X as:

{ }

Another useful parameter to describe the statistics of a RV X is variance.

The variance of a RV is a measure of the distance from the outcomes to the

11
expected value, more commonly known as dispersion. A small variance indicates

that the outcomes are very close to the expected value. In contrast, a large

variance indicates that the data is very spread out around the mean. Variance

can be estimated as:

{ } ∑ (∑ ) . (2.4)

2.3. Markov Process and Markov Chains

As we introduced in the previous section, a random or stochastic process

is a sequence of outcomes that depend on some probability. A Markov process is

a stochastic “memory-less” process in which predictions of future outcomes of

the stochastic process are based only on its present state. A random variable is

defined as a Markov process if the RV has the following properties:

a) The number of possible outcomes or states is finite;

b) The outcome at any stage depends only on the outcome of the previous

stage; and

c) The probabilities are constant over time.

Consequently a Markov process can be defined as follows:

( | ) ( | ) (2.5)

A sequence of random variables generated by a Markov process is called

a Markov Chain. Any Markov Chain is defined by the transition probability to

12
move from one stage to another. The probability that an RV in state xi will change

to a state xj in a single step is defined as:

( ) ( | ). (2.6)

The probability that the chain is in state xi at step n can be calculated

using the Chapman-Kolmogorov equation that is the sum of the probability of

being in a particular state at the current step and the transition probability from

the current state to xj:

( ) ∑ ( | ) ( ). (2.7)

A more convenient way to express the Chapman-Kolmogorov equation is

the matrix form:

( ) ( ) , (2.8)

where S(n) is the probability vector at the nth step, S(0) is the probability vector at

the initial step, and P is the probability transition matrix.

A Markov Chain can be illustrated using a simple example. Suppose that the

price variation of a particular share in the stock market can be predicted using a

Markov process and the possible price states are: “decrease”, “same” or

“increase”. Therefore, the price of the share tomorrow only depends on today’s

13
share price. Assuming that today the price variation is “same”, the transition

probability matrix (Figure 2.3) can be written as follows:

( ).

The initial state vector is:

( ) ( ).

Tomorrow’s price variation will be:

( ) ( )( ) ( ).

The sixth day’s price variation will be:

( ) ( )( ) ( )

14
Figure 2.3: Schematic representation of the transition probability matrix.

The summary of the Markov Chain for the first six days is shown in Table

2.3. However, if the initial price variation is “decrease” instead of “same” (see

Table 2.4), the Markov Chain reaches the same probability values at the sixth

day. Hence, on the sixth day, the Markov Chain reaches its stationary

distribution, where the probability values are independent of the initial state.

Table 2.3: Markov chain for initial price variation “same”.

Day Decrease Same Increase

1 0.2500 0.5000 0.2500

2 0.3000 0.4000 0.3000

3 0.3100 0.3800 0.3100

4 0.3120 0.3760 0.3120

5 0.3124 0.3752 0.3124

6 0.3125 0.3750 0.3125

15
Table 2.4: Markov chain for initial price variation “decrease”.

Day Decrease Same Increase

1 0.4000 0.3000 0.3000

2 0.3250 0.3600 0.3150

3 0.3145 0.3720 0.3135

4 0.3129 0.3744 0.3128

5 0.3126 0.3749 0.3126

6 0.3125 0.3750 0.3125

2.4. Markov Chain Monte Carlo Methods

Monte Carlo Simulation (MCS) methods are a set of computational

algorithms used to solve numerical problems based on random numbers. One of

the most common applications of MCS is to calculate the probability of a specific

outcome by randomly sampling a large number of scenarios. MCS methods

have many applications in physics, finance, engineering, and mathematics. In

this dissertation we use the MCS to explore the space of possible solutions used

in the proposed seismic inversion algorithm.

Markov Chain Monte Carlo (MCMC) is a family of algorithms for sampling

desired probability distributions based on Markov processes. The shape of the

desired distribution has to be similar to the MCMC equilibrium distribution; this

means that the quality of the MCMC distribution improves with the number of

steps. A Markov Chain is an ideal method for Monte Carlo simulations due to the

16
memoreless nature of Markov stochastic processes. This Monte Carlo method

combines all the desired features for stochastic simulation:

a) It is simple and easy to construct a Markov Chain with the desired

properties;

b) It generates statistically consistent samples from the target distribution;

and

c) The samples drawn are dependent and follow a Markovian framework

which allows more robust convergence.

Markov Chain mixing time is by definition the number of steps needed to

reach the equilibrium with tolerable error. The better the Morkov Chain, the faster

the mixing time. Figure 2.4 shows an example of a Markov Chain with a

Gaussian as steady distribution; in this example, the desired distribution is

reached after 50,000 steps.

17
500 Samples 1000 Samples
40 60

50
30
40

Counts
20 30

20
10
10

0 0
-4 -2 0 2 4 -4 -2 0 2 4

5000 Samples 10000 Samples


300 600

250 500

200 400
Counts

150 300

100 200

50 100

0 0
-4 -2 0 2 4 -4 -2 0 2 4

25000 Samples 50000 Samples


1500 4000

3000
1000
Counts

2000

500
1000

0 0
-4 -2 0 2 4 -4 -2 0 2 4
Property Property

Figure 2.4: Markov Chain evolution after 500, 1,000, 5,000, 10,000, 25,000 and
50,000 steps. The equilibrium distribution is a Gaussian with zero mean and
variance of one.

2.5. Metropolis–Hastings Algorithm

A random walk (RW) is a mathematical stochastic path that is generated

from a series of random steps. RW theory plays an important role in physics,

statistics and engineering, since many physical phenomena can be modeled

using a stochastic method; coin tosses, Brownian motion and stock prices are

examples of processes that can be modeled as RWs. These methods are

particularly useful as sampling technique since RW sampling methods are least

18
biased compared to systematic sampling algorithms. However, RWs can lead to

poor representation of the real distribution in very-large-dimension problems.

Figure 2.5 shows a two-dimensional RW with 5,000 steps. In a simple

RW, the location can only jump to neighboring sites of the lattice. In a simple

symmetric RW on a locally finite lattice (Figure 2.5), the probabilities of the

location jumping to each one of its immediate neighbors are the same.

Figure 2.5: Two-dimensional RW with 10,000 points.

RW algorithms are generally based on Markov Chain processes. In this

dissertation, we implemented a Markov Chain Monte Carlo method called the

Metropolis–Hastings algorithm. The Metropolis–Hastings algorithm was originally

introduced in 1953 by a group of scientists (Metropolis et. al., 1953) from Los

19
Alamos National Laboratory, and extended to its general form by W. K Hasting in

1970 (Hastings, 1970). The idea of the Metropolis–Hastings Monte Carlo

algorithm is to compute values of the function P(x) that are proportional to a

known probability distribution F(x) P(x)=K*F(x), where the proportional constant K

is unknown and very difficult to estimate. The Metropolis algorithm generates a

sequence of draws as follows:

1. Randomly select an initial value x0 that satisfied the relation F(X0)>0.

2. Using the current value (xt) draw, generate the candidate value (xt+1) from

a proposed transition probabilistic function q(xt,xt+1).

3. From the candidate (xt+1) and the current (xt) points, calculate the ratio of

the density at the candidate and current points:

( ) ( )
; (2.9)
( ) ( )

If α≥1 accept the candidate point (xt+1) as current value and return to step

2; otherwise, accept the candidate with probability α, else reject it and return to

step 2.

Finally, the Metropolis–Hastings algorithm can be written as follows:

( ) ( )
( ). (2.10)
( ) ( )

20
As with any MCMC methods, the Metropolis–Hastings algorithm

approaches the desired distribution as the number of steps increases. A key

issue in the successful implementation of the Metropolis-Hastings or any other

MCMC sampler is the number of runs (steps) until the chain approaches

stationarity (the length of the burn-in period). Typically, the first 1,000 to 5,000

elements are thrown out, and then one of the various convergence tests is used

to assess whether stationarity has indeed been reached.

A poor choice of starting values and/or proposal distribution can greatly

increase the required burn-in time, and an area of much current research is

whether an optimal starting point and proposal distribution can be found. One

suggestion for a starting value is to start the chain as close to the center of the

distribution as possible, for example by taking a value close to the distribution’s

mode. A chain is said to be poorly mixing if it stays in small regions of the

parameter space for long periods of time, as opposed to a well-mixed chain that

seems to effortlessly explore the space.

21
3. THE STATISTICAL ROCK PHYSICS MODEL

3.1.Introduction

This chapter introduces the concept of the statistical rock-physics model

and illustrates how well-log data and theoretical rock-physics models are used to

populate the model space of the proposed seismic-inversion algorithm.

Sedimentary rocks are aggregates of compacted crystalline grains and

organic matter that are often amalgamated by cement minerals. The free space

between the mineral grains is often called intergranular space or porosity. In in

situ conditions, the intergranular space is filled by liquid or gaseous fluids. The

solid portion of the rock is known as the rock’s matrix, or more commonly, “the

matrix”. The rock’s matrix geometry or how the crystalline grains are arranged in

a sedimentary rock is usually extremely complex and depends on many

variables, such as provenance, sorting, burial history, diagenesis, etc. We need

to define an enormous number of parameters to accurately relate the geometry

of most sedimentary rocks and their elastic moduli; among these parameters, we

have volume fractions of minerals, nature of boding on the grain interface, type

and volume fraction of cements, etc. (Ruiz, 2009).

3.2. Rock Physics Models

Rock-physics models are mathematical models that are related to rock

properties (e.g., porosity, mineral and cement composition pore geometry, elastic

22
moduli, etc.). Rock-physics models are used to estimate rock properties that

have not been sampled. Most rock-physics models can be classified into three

groups: theoretical, empirical, and heuristic (Avseth et al., 2005).

According to Avseth (2005), theoretical models are continuum mechanics

approximations of the elastic, viscoelastic, or poroelastic properties of rocks. Biot

theory (Biot, 1956) is one of the first to model the coupled mechanical behavior of

a porous rock filled with a linearly viscous fluid. At zero frequency, the Biot

relations turn into the Gassmann equation (Gassmann, 1951).

Empirical models are based on the statistical regression analysis of

experimental observations. Most empirical models are only valid for the specific

data sets and physical conditions where the data was measured. Empirical rock

physics models are widely used to predict the properties of rock for which

measured data is not available. Examples of empirical models are Castagna’s

mudrock line (Castagna et. al., 1985), Gardner’s density relation (Gardner et. al,

1974) and Han’s Gulf of Mexico relations (Han, 1986).

Heuristic rock-physics models are mathematical relations based on the

experience-based techniques for problem solving. Since these models are non-

mathematically rigorous, they are not optimal. Popular examples of heuristic

models are the Wyllie time-average equation (Wyllie et. al,. 1956) and the

Greenberg and Castagna relations (Greenberg and Castagna, 1992).

23
3.3. Deterministic Versus Statistical Models

A deterministic model is a mathematical relation in which outcomes are

precisely determined through known relationships among variables. In such

models, a given input will always produce the same output. In deterministic

models the estimation error is assumed to be zero or negligible. Unfortunately, in

Earth science it is impossible to gather the required knowledge to create a

deterministic model that accurately relates the physical properties of most

sedimentary rocks. In contrast, statistical models generate a set of possible

values with their associated probability of occurrence (Figure 3.1). Therefore,

theoretical and empirical rock physics models can be considered deterministic

and statistical models respectively. While deterministic models rely on the

physics of the phenomenon, probabilistic models trust in an adequate statistical

sampling of the analyzed set of data. It is important to emphasize that a poor

sampling of the data space will lead to the formulation of an inaccurate or

incorrect probabilistic model.

24
Figure 3.1: Deterministic (left) versus stochastic models (right). The stochastic
model generates a set of possible values instead of a simple outcome.

3.4. Probabilistic Rock-Physics Models

Probabilistic rock physics models are physical models that relate the rock

properties of a given seismic lithofacies using probabilistic techniques. Avseth

(2000) defined a seismic lithofacies as a seismic-scale sedimentary unit that is

characterized by its lithology, bedding configuration, petrography and elastic

properties. Seismic lithofacies can be used to estimate the statistical parameters

of the rocks that constitute the sedimentary unit. Since these rocks have similar

depositional and diagenesis history, we can assume the stationarity of the

statistics of the rock’s properties for any particular seismic lithofacies (Avseth,

2000).

25
We usually observe data scatter when we plot elastic properties against

rock properties for a given seismic lithofacies. Due to data scatter, descriptive

statistics are often used to summarize the elastic properties of a given seismic

lithofacies. The descriptive statistics is useful to define the most important

properties of the statistical populations. Figure 3.2 shows the P-wave velocity

histograms of two different lithofacies; the histograms clearly show that

lithofacies A has a lower most expecting value and larger velocity dispersion than

lithofacies B. Descriptive statistical measurements are normally used to define

the tendency of the data (Table 3.1), for example, when most of the data is

gathered or gives an idea of the data dispersion.

Figure 3.2: P-wave velocity histograms for two different lithofacies.

26
Table 3.1: Vp statistical summary table for seismic lithofacies A and B.

Seismic Lithofacies A Seismic Lithofacies B


Mean 3849.9 4553.2
Median 3831.4 4568.4
Standard Deviation 187.85 100.31
Variance 35288 10063

For most practical applications, descriptive statistics is not enough to

represent the complexity of the relations among rock properties, since the

definition of central estimators (like the mean or the median) and the dispersion

estimators (like the variance and standard deviation) generally do not have clear

meaning in multidimensional non-parametric spaces. Hence, to avoid the

limitations of descriptive statistics, non-parametric probability density functions

(PDFs) are introduced to better describe the dependence among rock properties,

including its non-uniqueness and non-linearity (Takahashi, 2000). Usually, PDFs

are generated from the integration of the available well and geological data and

are supported by theoretical rock physics models. In fact, multivariate PDFs are

usually hybrid rock physics models that can describe measured rock samples

and scenarios not sampled by measured data.

3.5. The multidimensional Non-Parametric Model

A poorly mixed chain can arise because the target distribution is complex

(multi-modal) and our choice of starting values traps the Markov Chain to one of

27
the modes. An approach to solve this problem is to transform the complex target

distribution to a simpler parametric distribution like a Gaussian distribution, and

then proceed with the MCMCS. This process is called normal scoring and is

commonly used in many statistical applications. Normal scoring processes can

be applied to any number of dimensions. However, it is easy to underestimate

the sampling (leaving many empty zones) of large-dimensional inverse problems.

To solve this issue, the number of samples has to be largely increased with the

number of dimensions. Figure 3.3 shows four three-dimensional projections of a

ten-dimensional Gaussian distribution sampled with 1,000, 10,000, 100,000 and

1,000,000 draws. We can see that the distribution sampled with 1,000 draws

does not represent the shape of the target Gaussian distribution. The modeled

distribution approaches the target distribution when the number of samples

exceeds 100,000 draws. After the MCMCS is performed, the resulting normal

Gaussian distribution is converted back to its original shape, applying reverse

normal scoring.

28
Figure 3.3: Three-dimensional projections of a ten-dimensional Gaussian PDF
sampled with 1,000 (upper left), 10,000 (upper right), 100,000 (lower left) and
1,000,000 (lower right).

3.6. Hybrid Rock-Physics Model Example

In hydrocarbon exploratory areas where the availability of well log data is

usually limited, rock physics modeling is used to populate the multivariate PDFs

with data or seismic lithofacies that are not sampled by the existing wells, but are

probable in the studied area (Gonzalez, 2006). An example of a hybrid rock

physics model can be made by the application of Gassmann’s equation

(Gassmann, 1951) to a set of well log data. Gassmann’s equation allows us to

29
estimate the effective bulk modulus of a fluid filled porous rock when the original

fluid is replaced by a second fluid. This is a very important problem in

hydrocarbon exploration, where fluid replacement is commonly used to sample

different fluids and saturations scenarios in potential reservoir rocks. Gassmann’s

equation is written as follows:

( )
(3.1)

, (3.2)

where:

= Effective bulk modulus of the rock with pore fluid,

= Effective bulk modulus of the dry rock,

= Bulk modulus of the pore grain material,

= Effective bulk modulus of the fluid,

= Porosity,

= Effective shear modulus of the rock with pore fluid,

= Effective shear modulus of the dry rock.

Some of the projections of a multivariable probability density function are

shown in Figure 3.4. The shown PDF was generated from well log data and it

defines the relations among the rock properties for clean sandstone. This

sandstone represents a lithofacies whose intergranular space is filled by brine.

30
Figure 3.4: 2D Projections of the multivariate probability distribution function. A)
Bivariate P-wave velocity versus S-wave velocity PDF; B) Bivariate P-wave
velocity versus bulk density PDF; C) Bivariate S-wave velocity versus bulk
density PDF; and D) Bivariate Porosity versus bulk density PDF.

We applied the Gassmann relations to predict the sandstone elastic

properties when the pores are filled with gas. We estimated the acoustic

properties of the brine and gas using the Batzle and Wang empirical model

(Batzle and Wang, 1992) and assuming salinity of 15,000 ppm, temperature of

84°C, pore pressure of 25 MPa and gas specific gravity of 0.6. The irreducible

water saturation of the modeled sandstone was assumed 8%; hence the fluid

properties were calculated for a saturation of 92% gas and 8% brine using the

31
homogeneous saturation model (Mavko et al., 1998). The fluid properties at the

specified temperature and pressure are displayed in Table 3.2.

Table 3.2: Fluid properties generated using Batzle and Wang (1992) fluid
properties empirical relations.

Brine Gas 90% Gas and 10% Brine

Bulk Modulus [GPa] 2.59 0.05 0.06

Density [g/cc] 0.99 0.126 0.21

Vp [km/sec] 1.61 0.63 0.53

The three-dimensional projection of the multivariate probability distribution

function of the modeled gas saturated sandstone is shown in Figure 3.5. As is

expected, the replacement of the original fluid with less dense and more

compressible fluid will decrease the density and bulk modulus. In contrast, the S-

wave velocity slightly increases due to the density effect on the S-wave velocity

expression. The comparison of the elastic properties histograms for the original

and fluid replaced sands shows that the fluid substitution introduces more

uncertainty in the gas filled rock in comparison with the brine-filled rock (see

Figure 3.6). A similar observation was made done by Sengupta (2000) in her

sensitivity analysis of fluid replacement.

32
Figure 3.5: 2D Projections of the modeled multivariate probability distribution
function. A) Bivariate P-wave velocity versus Vs PDF; B) Bivariate P-wave
velocity versus bulk density PDF; C) Bivariate S-wave velocity versus bulk
density PDF; and D) Bivariate Porosity versus bulk density PDF.

33
Figure 3.6: Elastic properties comparison between original saturation (left) and
gas replaced saturation (right).

This probabilistic rock physics modeling process can include any simple or

complex rock physics model as long as the rock property can be defined in the

multivariate PDF. For instance, the Dvorkin cemented sand model (Dvorkin and

Nur, 1996) can be used to theoretically add cement to sand measurements.

Other applications of theoretical rock physics models are shown in Figure 3.7.

The effect of fluid replacement, adding cement or reducing porosity over the P-

wave velocity (Vp) are displayed in the porosity-Vp space. Figure 3.8 shows the

probability functions for the modeled seismic lithofacies.

34
Figure 3.7: Example of theoretical rock physics models application to measured
P-wave velocity samples. The fluid replacement, cement modeling and porosity
perturbation effects are shown by red, green and black arrows.

Figure 3.8: Example of theoretical rock physics models application to measured


P-wave velocity samples. The probability distribution functions for the fluid
replacement, cement modeling and porosity perturbation effects are in red, green
and black, respectively.

35
3.7. Using MCMCS to Explore Model Space

As we have mentioned, our proposed stochastic inversion scheme relies

on a Markov Chain Monte Carlo Simulation (MCMCS) engine to explore the

model space represented by a non-parametric multivariate probability density

function. The use of Monte Carlo simulation is central to the resolution of inverse

problems with large-dimensional model spaces for which systematic evaluation

of model spaces on a regular grid is not feasible (since the required number of

points to evaluate is too large).

When the MCMCS algorithm is well implemented, the shape of the target

PDF does not change with the number of simulated samples; this means that the

MCMCS is really random and unbiased. An example of the MMCMCS

implementation is shown in Figure 3.9. The shape of the simulated distribution is

very similar to the original PDF, even though the number of simulated points is

more than one hundred times the number of original data points. We use the

Mean Squared Error (MSE) between the simulated and target distributions to

estimate the optimum number of steps to reach MCMCS equilibrium.

Figure 3.10 shows the MSE plot for the MCMCS example shown in Figure

3.9. For this particular MCMCS implementation the equilibrium is reached after

8,000 steps. The MSE is estimated as follows:

( ) ],

Where = Target Distribution and =Simulated Distribution.

36
Figure 3.9: Comparison of an original bivariate distribution (left) and the MCMCS
distribution (right).

Figure 3.10: Mean-square-error plot for the MCMCS of the example shown in
Figure 3.9.

37
4. 1D MARKOV CHAIN STOCHASTIC SEISMIC INVERSION

4.1. Introduction

This chapter presents the 1D Markov Chain stochastic inversion algorithm. First,

it shows the general workflow for the 1D inversion scheme. Second, it describes

in detail all the steps of the general 1D workflow.

In physics the inverse problem consists of using measurements to infer

the parameters that describe a system (Tarantola, 2005). In contrast, the forward

problem involves the prediction of the measurements given the parameters that

describe the studied system (Figure 4.1). In Earth science the goal of the inverse

problem is to estimate the earth’s properties from a series of geophysical

observations.

Figure 4.1: The inverse and forward problems are complementary mathematical
problems.

38
The content of this chapter was partially presented in the paper IPTC-18040 of the 2014 International Petroleum
Technology Conference in Kuala Lumpur.
This dissertation is focused on the specific problem of estimating the

earth’s parameters from surface seismic data. As with any inverse problem in

earth science, seismic inversion does not have a unique solution. Hence, we

proposed a novel inversion scheme that is based on a probabilistic method

instead of a deterministic approach. When the inverse problem is solved using a

probabilistic approach the system parameters are represented by a PDF

generated from the available data and model. The probabilistic theory applied to

this dissertation can be applied in any kind of inverse problem, including

nonlinear problems. To solve most probabilistic inverse problems, we need to

implement an extensive exploration of the system space. Due to the

multidimensional nature of the Earth science problems, we cannot perform a

systematic system space exploration. Instead, we perform possible solution

space exploration using Markov Chain Monte Carlo methods.

4.2. Proposed Inversion Algorithm

The proposed 1D stochastic inversion algorithm is based on a Markov

Chain stochastic simulation that uses statistical rock physics to constrain the

space of possible solutions. Mukerji et al. (2001) introduced the statistical rock-

physics methods as a way to combine rock physics, information theory, and

geostatistics to reduce uncertainty in reservoir property estimation. Our

fundamental contribution is using the Metropolis–Hastings MCMCS algorithm to

generate multivariate non-parametric statistical rock physics models and

introducing the inversion technique that directly produces rock property volumes

39
instead of elastic property volumes; using our methods, inversion results can be

directly introduced into the geological model of the studied area.

Figure 4.2 shows an idealized scheme for a 1D inversion algorithm. The

inversion technique consists of four main steps: the rock properties scenario

generation, the synthetic data calculation and the comparison between the real

and synthetic traces. In each realization, a model of the subsurface properties is

generated, and the model properties are calculated from the previously defined

multivariate PDF. After the subsurface model is generated, a synthetic trace is

produced using the elastic properties from the subsurface model. Finally, the

synthetic trace is compared to the real seismic trace; if the similarity between

both traces is larger than a given threshold, the model is accepted as a possible

solution and stored. In contrast, if the likeness between the real and the synthetic

traces is smaller than the mentioned threshold, the model is erased and a new

model is generated.

An important characteristic of this proposed inversion scheme is that the

model is generated in a depth domain, so there is no need to convert the model

to depth after a probable subsurface model has been accepted. A detailed

description of each inversion step is shown in the Chapter 5. The proposed

algorithm produces rock property traces instead of elastic property traces.

Therefore, the inversion outcome can be directly introduced into the geological

model of the studied area.

40
Figure 4.2: Simplified 1D stochastic inversion workflow.

4.3. Possible Scenario Generation

Every 1D model is represented by two main attributes. The first is the

layers geometry or how the seismic lithofacies are arranged within the studied

interval. Each seismic lithofacies is defined as a discrete variable with a related

rock property distribution. The second is the rock properties of the analyzed

lithofacies. The scenario generator algorithm works on two levels. First the

routine generates the geometry of the probable earth models. Once the geometry

41
is built, we assign rock properties to each layer. The rock properties are assigned

by using a Markov Chain-based stochastic simulation algorithm. This process is

illustrated in Figure 4.3. First, the fraction of the seismic lithofacies that constitute

the analyzed interval is generated. Once the 1D array with the fractions of the

seismic lithofacies that constitute the study interval are generated, the 1D array

with the distribution of these lithofacies is produced using the Fisher-Yates

shuffle algorithm. This algorithm was originally described by Ronald A. Fisher

and Frank Yates in 1938 (Fisher et al., 1938). The Fisher-Yates shuffle algorithm

is a numerical method for generating random permutations of a finite number of

groups. The most important property of the Fisher–Yates shuffle is that it is

unbiased; consequently, every permutation is equally probable. The algorithm to

generate a random permutation of the numbers one through N goes as follows:

a) Write down the numbers from one through N;

b) Pick a random number X between 1 and the number of remaining

unchosen numbers;

c) Counting from the low end, strike out the Xth number not yet selected and

store it;

d) Repeat from step 2 until all the numbers have been selected.

The 1D array of numbers stored after all the numbers are selected is now

a random permutation of the original numbers. If the random number generator is

truly random and unbiased, the resulting permutation will be also.

42
An example of the application of the Fisher and Yates method, for an

interval that ranges from 1 to 5, is illustrated in Table 4.1.

Figure 4.3: Schematic representation of the possible scenario generation.

Table 4.1: Example of the Fisher and Yates shuffle algorithm. The permuted
interval ranges from 1 to 5. The red cells correspond to the number that is
selected in each iteration.

Step Random Number Sequence Result


1 4 1 2 3 4 5 4
2 1 1 2 3 4 5 41
3 3 1 2 3 4 5 415
4 2 1 2 3 4 5 4153

4 1 1 2 3 4 5 41532

43
The generated seismic lithofacies indicator 1D array is the template used

for the rock property 1D arrays generation. The property 1D arrays are produced

in an order that starts with the lithology and porosity, and finishes with the elastic

properties. Each cell in the property 1D array is filled with the property

information calculated from the multivariate PDF and defined by the Markov

Chain. The accuracy of the rock properties model relies on the idea that the

multivariable PDF truly represents the statistics of the target lithofacies.

We have also included some additional constraints for our inversion:

a) Fluid contacts: For geological models with local fluid contacts, all samples
of gas reservoirs should be above the oil samples and all samples with oil
should be above the water samples.

b) Geological age: The seismic lithofacies geometry can be constrained in a


given order that follows a stratigraphic column.

c) Layer thickness: The thickness of the layers can be constrained to any


given range.

4.4. Synthetic Seismic Generation

The probabilistic inverse problem defined by Equation 1.1 estimates the a

posteriori PDF ( ( )) from the a priori PDF of the data ( ( )) and the model

( ( )). Equation 1.1 is valid when the information in the space of measured

parameters (data) has been obtained independently of the prior information in the

model space (Tarantola, 2005). In section 1.1, we introduced forward modeling

operator ( ( )) that predicts data (Equation 1.2) from the model. The predicted

44
values are not necessary identical to the observations since there are

uncertainties related to the measurements and the proposed model.

In the present study the forward modeling operator is implemented as a

convolutional model for the synthetic trace generation. This model is expressed

as follows:

( ) ( ) ( ), (4.1)

where:

( )= Synthetic seismic trace,

( )= Angle dependent reflection coefficients in time domain,

( )= Angle dependent extracted wavelet, and

‘*’ = Convolution operator.

The convolutional model of the seismic trace assumes that the latter is the

superposition of the reflection coefficients filtered by the bandwidth of the

recorded wavelet (Sheriff, 1989). Since the elastic property 1D arrays are

generated in the depth domain, they have to be stretched to two-way time using

a depth-to-time function. We assumed here that the depth-to-time conversion

function is part of the input data. Another important part of the input data is the

set of seismic wavelets. Seismic wavelets have to be previously extracted from

the measured seismic data at the same angle ranges at which the synthetic data

have been generated.

45
After the elastic properties arrays are converted to two-way travel time,

they are interpolated to the same regular sampling as the input seismic data.

This is a very important process since the time sampling acts as a frequency

operator that defines the maximum vertical resolution of the inverted earth model

(see Figure 4.4). Finally, a Nyquist filter is applied to the time-stretched elastic

models to avoid any possible aliasing in the time domain.

Figure 4.4: Comparison among the depth, time and regularized time 1-D P-wave
velocity arrays. The black circles show how the regularization in the time domain
can affect the elastic properties of the modeled subsurface.

Once the elastic properties arrays in time have been generated, the angle-

dependent reflectivity profiles are produced using the Aki and Richards linear

approximation (Aki and Richards, 1980) of the Zoeppritz equation. The Aki and

46
Richards linear approximation can be expressed in terms of P-wave velocity, S-

wave velocity and bulk density (ρ) as follows:

( ) ( ) [ ( )] ( ) (4.1)

where:

ΔVp=Vp2-Vp1 Vp=0.5(Vp1+Vp2)

ΔVs=Vs2-Vs1 Vs=0.5(Vs1+Vs2)

Δρ=ρ2-ρ1 ρ=0.5(ρ1+ρ2)

ϴ=0.5(ϴ1+ ϴ2);

Vp, Vs, ρ and ϴ are the P-wave velocity, S-wave velocity, bulk density and

incidence angle, respectively. The sub-indexes 1 and 2 refer to the upper and

lower layers correspondingly. Equation 4.1 can be rewritten as:

( ) ( ) (4.2)

where A, B and C are the AVO intercept, AVO gradient and AVO curvature. We

used the first two terms of AVO Equation 4.2 for the estimation of the angle-

dependent reflectivity arrays. The two-term Aki and Richards AVO equation is

valid for a maximum incident angle of about 36°.

Finally, the synthetic trace is calculated by the convolution of angle-dependent

reflectivity converted to two-way time and the angle-dependent wavelet. Figure

4.5 shows a schematic representation of the synthetic traces generation step.

47
Figure 4.5: Schematic representation of the synthetic traces generation
algorithm.

4.5. The Misfit Function

The final step in the 1D stochastic inversion scheme is the estimation of

the misfit function between the real and synthetic seismic traces. The synthetic

trace is compared to the real seismic trace and if the difference between them is

less than a given threshold the model is accepted; otherwise, the model is

erased and a new one is generated.

In the proposed algorithm, the similarity analysis is performed in a

cascade sequence that starts with the nearest and finished, with the farthest

angle-stacked traces. Once a trace is accepted the algorithm analyzes the next

angle trace until all the angle-dependent reflectivity traces are accepted. There

are many criteria to estimate the degree of similarity between the synthetic and

48
real traces; we used L2-norm to measure the similarity between the real and

synthetic traces. The L2-norm can be expressed as follows:

√∑ ( ) , (4.2)

where: Ti Syn= Synthetic trace, Ti Real= Real trace, and i= ith Element of the

1-D trace array.

An example of one trace misfit analysis is shown in Figure 4.6. The top

graph shows the real (black) and synthetic (blue) traces. The difference between

the synthetic and real traces is displayed in the center graph. Finally, the

cumulative L2-norm is plotted in the bottom graph. In our proposed algorithm, we

select the models that minimize the L2-norm.

Figure 4.6: Misfit calculation example. In this synthetic example the cumulative
2
L -norm shows the intervals with larger misfit.

49
The probable models generated by the proposed inversion scheme are a

valid solution to the studied inverse problem; all of them honor the input rock

physics model as well as the measured seismic data.

50
5. SYNTHETIC EXAMPLES

5.1. Introduction

The central problem in seismic reservoir characterization is the prediction

of lithology and saturation from seismic and well log data. In this chapter we

present the results of the application of our proposed inversion scheme on two

different sets of synthetic data. The synthetic data sets were generated based on

real well data with the goal of testing the inversion in two different geological

environments. The first synthetic data set is composed of low porosity high

impedance sand. The goal of the second example is to replicate a high porosity

low impedance clastic environment. In this example, the target reservoir rock is

characterized by weakly compacted uncemented sandstones.

In both synthetic tests, the model data space is represented by the

multivariate PDFs that were generated from the synthetic log and extended by

theoretical models. For instance, changes in pore fluid properties were modeled

using Gassmann fluid replacement theory. The rock’s properties in both synthetic

log data sets were based on real measured data. Next, the data space was

generated using synthetic seismic data modeled from synthetic elastic well logs.

As discussed in Section 4.4, the forward modeling operator was based on the

convolutional model of the seismic trace.

51
The content of this chapter was partially presented in the paper IPTC-18040 of the 2014 International Petroleum
Technology Conference in Kuala Lumpur.
5.2. Synthetic Test 1: Cemented Sand

The first synthetic data set was based on real well log data measured in a

Cretaceous interval; the in situ well logs are displayed in Figure 5.1. The

synthetic model was formed by two sands encapsulated into a shale sequence.

The top upper sand is located at 3,347 meters depth with a thickness of 40

meters. The lower sand is thinner with a thickness of 20 meters and located just

32 meter bellow the base of the upper sand. The target reservoirs are cemented

sand with average porosity of 16%. The background P-wave impedance is about

10 km*kg/sec*cc. The average P-impedance corroborates that the sands are well

compacted and cemented.

Figure 5.1: Model 1 synthetic well logs showing (from left to right) facies
indicator, shale volume, P-wave velocity, bulk density, S-wave velocity, porosity,
P-wave impedance and S-wave impedance.

52
In this kind of elastic scenario, fluid variations have very small influence on

the elastic properties of the target reservoirs. In high-impedance reservoir

settings, the goal of quantitative interpretation is usually the prediction of gross

reservoir thickness and reservoir quality. In our study here, the gross reservoir

thickness is expressed by the fraction of the potential reservoir rock in the

inverted interval. The reservoir quality is defined by two properties, which are the

fraction of shale (shale volume) and the total porosity of the sandstone. The

study reservoir properties are described by non-parametric histograms and

descriptive statistical estimators.

The elastic fluid properties profiles for brine, oil and gas were estimated

using the Batzle and Wang (1992) equations and assuming pressure and

temperature gradients of 9.72 MPa/km and 22.3 °C/km. Oil gravity, gas gravity

and salinity were 36° API, 0.6 and 8,000 ppm respectively. Figure 5.2 shows the

P-wave and density profiles for three possible pore fluids. Once the fluid

properties are estimated, the fluid replacement is performed using the Gassmann

relations. The fluid substitution was performed from brine to 90% hydrocarbon

saturated sands.

53
Figure 5.2: Synthetic model 1 pore fluids velocity and density profiles.

The final seismic lithofacies multivariate PDF projections for the Vp-Vs,

Vp-Rhob, and PhiT-Vp spaces are shown in Figure 5.3. The effect of the fluid

variations on the elastic properties of the permeable rocks is negligible. However,

there was good lithology discrimination in all projected spaces. Hence, the test

was planned focus on the prediction of gross reservoir thickness and reservoir

quality.

54
Figure 5.3: Final seismic lithofacies multivariate PDF projections for the Vp-Vs
(Upper), Vp-Rhob (Center), and PhiT-Vp (Lower) spaces. Green, blue, brown
and red contours denote shales, brine sands, oil sands and gas sands,
correspondingly. Some empirical models are plotted for comparison including
Castagna’s mudrock line, Castagna’s 1993, Han’s 1986, Gardner’s sand,
Gardner’s shale, Voigt bound, Reuss bound and Nu’s modified Voigt bound for
quartz mixture.

The angle-dependent reflectivity curves were generated from the synthetic

logs converted to time. They showed strong positive reflection coefficients at the

top of both sands. It was clearly evident that one of the challenges in this test

was to accurately predict the reservoir geometry in the target interval. Figure 5.4

55
shows the model 1 angle dependent reflectivity profiles. It can be observed how

the reflectivity of the top and base of the sands decrease with the angle of

incidence. This is typical reflectivity response of high impedance sands.

Figure 5.4: Model 1 angle dependent reflectivity profiles.

Next, we generated the synthetic angle stack traces by convolving the

angle-dependent reflectivity with a Ricker wavelet with a central frequency of 25

Hz (Figure 5.5).

56
Figure 5.5: Ricker wavelet with a central frequency of 25 Hz.

For this test we assumed the stationarity of the wavelet, meaning that

wavelet did not vary with incident angle or propagation time. The results of this

forward modeling are shown in Figure 5.6. A set of five angle traces were

generated equally spaced in an interval that ranged from 0° to 32°. This angle

gather was used as the measured seismic data for this synthetic test.

Figure 5.6: Synthetic angle gather generated from the synthetic model 1.

57
Once the “real traces” were generated the next step was to produce the

multivariate PDF for each of the analyzed lithofacies. The 2D projections of the

multivariate PDF (Figure 5.6) show that the sands have larger p-wave velocity,

lower Vp/Vs ratio and larger density than the surrounding shales. The

observations in Figure 5.7 indicate that there is an excellent lithofacies

discrimination based on the elastic properties of the analyzed statistical

populations.

Figure 5.7: 2D Projections of the multivariate probability distribution function for


sands (left) and shales (right) calculated from the synthetic model 1.

58
The inversion was performed for one million realizations using the model

shown in Figure 5.6 as the measured data and the multivariate PDFs shown in

figure 5.7 as the rock physics models. Modeled traces with similarity factor (α)

smaller than 0.9 were selected as possible solutions for the stochastic inversion

of the synthetic data. The comparison between the synthetic seismic data and

the 1,112 possible solutions revealed the close similarity between the synthetic

measured data and the probable models seismic traces. The residuals were

small, the time arrivals (Figures 5.8) and waveforms (Figures 5.9) similar.

Figure 5.8: Real seismic trace (left) and possible models traces (right)
2
comparison. The selected traces have L -norms smaller than 0.9.

59
Figure 5.9: Synthetic versus modeled waveform comparison for four possible
models with α<0.9.

The seismic lithofacies possible models with α<0.9 are displayed in Figure

5.10. It can be observed that most of the sand samples are located inside the

limits of the upper and lower sands. In fact, these sands are well-delimited by

most of the probable inversion solutions. As discussed Section 5.1 we are

interested in the prediction of gross reservoir thickness and reservoir quality.

Gross reservoir thickness is defined by the fraction of sand in the target interval.

Figure 5.11 shows the PDF of the sand fraction for all probable models with

60
α<0.9. The most expecting value of the sand fraction is 0.13, which is the same

fraction defined in the synthetic model. However, there is not a sole solution,

since the inversion produces a range of likely solutions that are constrained by

the rock-physics model.

Figure 5.10: Probable models with α<0.9. The black, red, yellow and green lines
represent the top of upper sand, base of upper sand, top of lower sand and base
of lower sand, respectively.

61
Figure 5.11: Sandstone fraction for probable models with α< 0.9.

Once the geometrical distribution of the two seismic lithofacies was

calculated from the probable models, we were able to define the key rock

properties for each lithofacies, and especially for the sandstones. For this

particular case, total porosity was a very important property for the definition of

reservoir quality. From the total-porosity models in Figure 5.12 we were able to

calculate some descriptive statistical coefficients; for this particular test, the most

likely sand total porosity was about 0.15, with a standard deviation of 0.7. From

the total porosity and volume of shale probable models (Figure 5.13), it was

possible to estimate the effective porosity of the target reservoir (Figure 5.14).

Finally, the elastic property solutions are shown in Figure 5.15.

62
Figure 5.12: Total porosity of the probable models with α<0.9. The black, red,
yellow and green lines represent the top of upper sand, base of upper sand, top
of lower sand and base of lower sand, respectively.

Figure 5.13: Volume of shale probable models with α<0.9. The black, red, yellow
and green lines represent the top of upper sand, base of upper sand, top of lower
sand and base of lower sand, respectively.

63
Figure 5.14: Effective porosity probable models with α<0.9. The black, red,
yellow and green lines represent the top of upper sand, base of upper sand, top
of lower sand and base of lower sand, respectively.

Figure 5.15: Elastic models with α<0.9. The black, red, yellow and green lines
represent the top of upper sand, base of upper sand, top of lower sand and base
of lower sand respectively.

64
As has been mentioned before, all the solutions generated by the

proposed seismic inversion scheme are probable answers to the analyzed

seismic inverse problem. Therefore, the statistics of the solutions have more

significant relevance than the models themselves. The comparison between the

rock properties of the synthetic model and the solutions (Figure 5.16) shows a

very close similarity between the synthetic data and the inversion solutions. The

synthetic model used in this exercise is realistic and it is based on measured well

log information.

Figure 5.16: Normalized histograms comparison between the rock properties of


the synthetic model and the probable solutions, for sand’s shale volume (upper
left), sand’s total porosity (lower left), shale’s shale volume (upper right) and
shale’s total porosity (lower right). Red and blue bars correspond to synthetic
model and solutions, respectively.

65
It is clear that if α decreases the similarity of the possible models will

increase. However, the number of possible models will decline as well. Figure

5.17 shows the possible solution models for α<0.045; in this particular case, we

obtained 78 possible models instead of the 1,112 calculated with α<0.9. This is

an important characteristic of the proposed inversion scheme; this is further

discussed in Section 7.4.

Figure 5.17: Probable models with α<0.045. The black, red, yellow and green
lines represent the top of upper sand, base of upper sand, top of lower sand and
base of lower sand, respectively.

5.3. Synthetic Test 2: High Porosity Sands

The first synthetic test proved the viability of the application of the

proposed inversion scheme to predict the geometry and quality of the reservoir of

66
the target formation. The goal of the second synthetic test is to replicate a high-

porosity, low-impedance clastic environment. In this example, the target reservoir

rock was characterized by uncemented sandstone with an average porosity of

31%. In this geological setting, fluid properties variations have very strong effect

on the elastic properties of the studied sands. The associated hydrocarbon

seismic signatures of these high-porosity reservoirs are usually anomalous high

amplitude reflections called “bright spots”. In this test, the objective was to define

the fraction of sand, the fraction of hydrocarbon sand, the fluid contact, the

volume of shale and the effective porosity of the analyzed sand.

The synthetic well logs, displayed in Figure 5.18, show the in situ case in

black line, brine case in blue and the gas case in red. The fluid replacement was

performed using the Gassmann relations. The elastic fluid properties profiles for

brine, oil and gas (Figure 5.19) were estimated using the Batzle and Wang

(1992) equations and assumed pressure and temperature gradients of 9.72

MPa/Km and 20 °C/Km. The gas gravity and salinity were 0.6 and 10,000 ppm

respectively. The well logs indicated that the brine-saturated sandstone had

larger impedance than the surrounding shale. However, when the pore fluid was

substituted from brine to a mixture of 90% gas and 10% brine, the sandstone’s

impedance became lesser than the shale’s acoustic impedance. Again, the rock-

physics diagnostic plot (Figure 5.20) shows the most important rock properties

for the three analyzed seismic lithofacies; in general, the elastic discrimination

among the seismic groups was good.

67
1950

2000

2050

2100

2150
Depth [m]

2200

2250

2300

2350

2400

2450
0 0.5 1 0 0.2 0.4 1.5 2 2.5 1.5 2 2.5 0.5 1 1.5
Vshale [fraction] Porosity [fraction] Vp [km/sec] Rhob [g/cc] Vs [km/sec]

Figure 5.18: Model 2 synthetic well logs showing (from left to right) facie
indicator, shale volume, porosity, P-wave velocity, bulk density and S-wave
velocity. Black, blue and red lines correspond to in situ, brine and gas
saturations.

1950

2000

2050

2100

2150
DEPTH [m]

2200

2250

2300

2350

2400

2450
0.5 1 1.5 2 0 0.5 1 1.5
Vp [km/sec] Rhob [g/cc]

Figure 5.19: Synthetic model 2 pore fluids velocity (left) and density profiles
(right). Blue and red lines correspond to brine and gas, respectively.

68
Figure 5.20: Final seismic lithofacies multivariate PDF projections for the Vp-Vs
(Upper), Vp-Rhob (Center), and PhiT-Vp (Lower) spaces. Grey, blue, and red
contours denote shales, brine sands and gas sands, correspondingly. Some
empirical models are plotted for comparison, including Castagna’s Mudrock line,
Castagna’s 1993, Han’s 1986, Gardner’s sand, Gardner’s shale, Voigt bound,
Reuss bound and Nu’s modified Voigt bound for quartz mixture.

The angle-dependent reflectivity plot (Figure 5.21) shows the main

impedance contrasts in the studied interval; the top of the gas sand is delineated

by a negative reflection coefficient with strong negative gradient. The largest

positive reflection coefficient corresponds to the gas-brine fluid contact within the

sandstone. This seismic interface has a relative small positive AVO gradient.

69
Finally, the base of the brine-saturated sand is defined by a negative reflectivity

with a strong negative gradient.

Figure 5.21: Model 2 angle-dependent reflectivity profiles.

The synthetic angle gather (Figure 5.22) was generated by convolving a

25 Hz Ricker wavelet (Figure 5.5) with the reflectivity logs displayed in Figure

5.22. As was expected, the most important reflection is the gas-brine fluid contact

situated at approximately 2.200 seconds. Next, the multivariate probability

functions were generated for the shale, the brine sand and the gas sandstone

seismic lithofacies (Figure 5.23).

70
Figure 5.22: Synthetic angle gather generated from synthetic model 2.

Figure 5.23: 2D Projections of the multivariate probability distribution function for


gas sands (left), shales (center) and brine sands (right), calculated from synthetic
model 2.

71
The inversion was five million realizations with the constraint that all brine

samples should be above the deeper gas sample. This is important because we

assumed a unique fluid contact in the whole interval; the models with L2-norms

smaller than 0.88 were selected as possible solutions for the inversion. The

comparison between the synthetic model and the 664 obtained solutions (Figure

5.24) shows a close similarity between the synthetic input data and the inversion

results. The waveform comparison for four inverted solutions and the synthetic

model (Figure 5.25) demonstrate that the main events are well reproduced by the

inversion algorithm.

0.08
2

0.06
2.05

2.1 0.04

Amplitude [Dimensionless]
2.15 0.02
Time [sec]

2.2 0

2.25 -0.02

2.3 -0.04

2.35 -0.06

2.4 -0.08

100 200 300 400 500 600


Number of Traces

Figure 5.24: Real seismic trace (left) and possible models traces (right)
2
comparison. The selected traces have L -norms smaller than 0.88.

72
Figure 5.25: Synthetic versus modeled waveform comparison for four possible
models with α<0.88.

The seismic lithofacies distributions for the 664 solutions are shown in

Figure 5.26. A visual inspection of the seismic lithofacies distributions shows a

large consistency for the depth of the top of the sand, fluid contact and base of

the sand. We generated histograms of the depth of the three seismic lithofacies

interfaces to evaluate the accuracy of the proposed inversion scheme (Figure

5.27).

Table 5.1 shows the comparison between the synthetic data and the

depths obtained by the inversion; these results demonstrate that the inversion

scheme is very accurate.

73
Figure 5.26: Probable models with α<0.88. The green, black and white represent
the shales, gas sands and brine sandstone seismic lithofacies, respectively.

400 400 350

350 350
300

300 300
250

250 250
200
Counts

Counts

Counts

200 200

150
150 150

100
100 100

50
50 50

0 0 0
2155 2160 2165 2170 2175 2190 2195 2200 2205 2210 2210 2220 2230 2240
Depth of Top of the Sand [m] Depth of Gas-Water Contact [m] Depth of Base of the Sand [m]

Figure 5.27: Depth of top of the sand (left), gas-water contact (center) and base
of the sand (right) histograms.

74
Table 5.1: Comparison between the synthetic data and the depths obtained by
the inversion.

Gas-Water
Top of Sand Base of Sand
Contact
Synthetic Model
2166.00 2202.00 2222.00
Value [m]
Mean [m] 2165.80 2198.70 2222.30
Median [m] 2166.00 2198.00 2222.00
Standard Deviation 2.82 2.63 3.21
Variance 7.97 6.92 10.34

Figures 5.28 to 5.31 show total porosity, shale volume, and effective

porosity, respectively. The effective porosity was calculated from the total

porosity and volume of shale properties. The elastic property solutions are shown

in figure 5.32. The fluid contact is easily observed in all the elastic models.

Total Porosity Vertical Models

2000 0.32

2050 0.3

Total Porosity [fraction]


2100 0.28

2150 0.26
Depth [m]

2200 0.24

2250 0.22

2300
0.2
2350
0.18
2400
0.16
100 200 300 400 500 600
Number of Traces

Figure 5.28: Total porosity of the probable models with α<0.88.

75
Volume of Shale Vertical Models

0.9
2000
0.8
2050
0.7

Shale Volume [fraction]


2100
0.6
2150
Depth [m]

0.5
2200
0.4
2250
0.3
2300
0.2
2350

2400 0.1

100 200 300 400 500 600


Number of Traces

Figure 5.29: Volume of shale probable models with α<0.88.

Effective Porosity Vertical Models

0.3
2000

2050 0.25

Effective Porosity [fraction]


2100
0.2
2150
Depth [m]

2200 0.15

2250
0.1
2300

2350 0.05

2400
0
100 200 300 400 500 600
Number of Traces

Figure 5.30: Effective porosity probable models with α<0.88 calculated from the
total porosity and shale volume properties.

76
Figure 5.31: Elastic models with α<0.88.

5.4. Conclusions

The synthetic test results presented in this chapter demonstrate the

feasibility of the estimation of rock properties by our proposed probabilistic

inversion approach. The first example showed the capability of the novel

77
inversion scheme to estimate two critical rock properties for hydrocarbon

exploration: porosity and reservoir fraction.

The second synthetic test included the fluid prediction as an additional

rock property to be estimated by the inversion. The outcome demonstrated that

the gas-filled sandstone can be easily discriminated from the shales and brine-

saturated seismic lithofacies. We also showed the accuracy of the method in the

estimation of the depth of the seismic lithofacies interfaces.

In conclusion, the results of these synthetic tests proved the feasibility of

using the proposed stochastic inversion technique to calculate the properties of

seismic lithofacies based on multivariate probabilistic rock physics models,

reflection seismic data, using Markov Chain Monte Carlos Simulation (MCMCS)

to sample the model space.

78
6. 1D REAL DATA APPLICATION

6.1. Introduction

This chapter presents the application of our proposed seismic stochastic

inversion algorithm to a real data set. The data set includes a group of well logs

and a 2D line of prestack seismic data extracted from a 3D volume crossing the

exploratory well. The studied stratigraphic sequence corresponds to tertiary

marine sediments deposited on a platform margin. From well-log data and the

geological information, we identified two main lithologies: sandstones and shales.

The potential reservoirs are unconsolidated sandstones with total porosity that

ranges from 35% to 39%.The pore space can be filled by gas, low saturation gas

(LSGS) or brine. Each combination of fluid and lithology is considered an

independent seismic lithofacies. Therefore, the target interval is made up of four

seismic lithofacies: shales, gas-filled sands, low saturation gas sands (LSGS)

and brine sands. Figure 6.1 shows the in situ saturation logs for the two hundred

and twenty meter study interval. The target interval has two sands: the upper

sand is interceded with total thickness of twenty meters and a sand fraction of

73%; the lower reservoir is eighty meters of blocky sand with three fluid faces:

gas, LSGS and brine.

The seismic data included raw PSTM gathers with a time sampling of 2

msec. The near angle stack generated from 0° to 15° is shown in Figure 6.2; the

top of the strong seismic anomaly related to the top of the gas reservoir is

79
represented by large negative seismic amplitude. The top of the upper and lower

sands are located at 72 ms and 104 ms, respectively. The goal of this inversion

was to predict total porosity, volume of shale, effective porosity, and fluid type

distribution in the study interval.

Figure 6.1: Well 1 logs for in situ saturation.

Figure 6.2: Near angle stack; the top of the upper and lower sands are located at
72 ms and 104 ms, respectively.

80
6.2. Seismic Data Conditioning

The prestack seismic data were conditioned to remove residual coherent

noise and to improve the signal-to-noise ratio for the stochastic seismic inversion.

This was an important stage, since the forward modeling operator assumed that

all the reflections corresponded to primary reflections and random noise, and no

coherent noise was present in the seismic data. The seismic data was also

resampled to one millisecond to match the time sampling of the synthetic models.

The angle gather at Well-1 location is displayed in Figure 6.3; the amplitude

extraction at the top of the upper sand shows a strong AVO effect associated

with the top of the gas sand.

Figure 6.3: Angle gather at the Well-1 location (Upper). The amplitude extraction
at the top of the upper sand shows a strong AVO effect (Lower).

81
After the seismic data was conditioned and converted to angle gathers,

the well logs were converted to time and the well-to-seismic tie was performed

(Figure 6.4). The well log tie shows close similarity between the zero offset trace

and the acoustic synthetic trace (right track). The angle-dependent extracted

wavelets are shown in Figure 6.5. The L2-norm calculated from the real and

synthetic traces is α=0.88. Therefore, we assumed that any inversion solution

with α≤0.88 was a possible solution to the studied real data application.

Figure 6.4: Well logs converted to time showing the well to seismic tie (right
track); the black line corresponds to the measured seismic trace and blue the
synthetic seismogram.

82
1

oo
0.8
8o

16o
0.6
24o
Amplitude [Dimensionless]

32o
0.4

0.2

-0.2

-0.4
-100 -80 -60 -40 -20 0 20 40 60 80 100
Time [ms]

Figure 6.5: Angle-dependent extracted wavelets.

6.3. The Model Space Generation

The available well-log information was used to generate the multivariate

PDF for the four observed seismic lithofacies: shale, gas sand, low gas

saturation sand and brine sand. Fluid replacement based on Gassmann’s

relation was performed to better understand the effect of fluid variations in the

target reservoir sands. Figure 6.6 shows the well logs for the in situ, 90% gas,

and 100% brine-saturated sands.

The 2D projections of the multivariate PDF (Figure 6.7) show the feasibility

of discriminating most of the seismic lithofacies groups based on their elastic

properties. Only the low saturation gas sands cannot be easily discriminated from

gas sands.

83
Figure 6.6: Well 1 logs after fluid substitution. Black, red and blue lines
correspond to in situ, 90% gas and 100% brine-saturated sands.

Shale Brine Sand Gas Sand LSGS


1.4 1.4 1.4 1.4
Vs [km/sec]

Vs [km/sec]

Vs [km/sec]

Vs [km/sec]
1.2 1.2 1.2 1.2

1 1 1 1
2 2.5 3 2 2.5 3 2 2.5 3 2 2.5 3
Vp [km/sec] Vp [km/sec] Vp [km/sec] Vp [km/sec]

2.2 2.2 2.2 2.2


Rhob [g/cc]

Rhob [g/cc]

Rhob [g/cc]

Rhob [g/cc]

2 2 2 2

1.8 1.8 1.8 1.8

2 2.5 3 2 2.5 3 2 2.5 3 2 2.5 3


Vp [km/sec] Vp [km/sec] Vp [km/sec] Vp [km/sec]
3 3 3 3
Vp [km/sec]

Vp [km/sec]

Vp [km/sec]

Vp [km/sec]

2.5 2.5 2.5 2.5

2 2 2 2

0.2 0.3 0.4 0.2 0.3 0.4 0.2 0.3 0.4 0.2 0.3 0.4
PhiT [fraction] PhiT [fraction] PhiT [fraction] PhiT [fraction]

Figure 6.7: 2D Projections of the multivariate probability distribution function


generated from Well 1 log data and Gassmann fluid substitution.

84
6.4. The Seismic Stochastic Inversion

The rock property models were generated assuming that all gas samples

were located above the low saturation gas sands and all LSGS were situated

above the brine sand. The fluid contacts can be situated at any depth as long as

the order of the fluid interfaces is followed. The properties of the seismic

lithofacies were assigned from the multivariate PDFs by using the MCMCS

engine to sample them. The proposed inversion algorithm generated data in two

domains; the first is the rock physics simulation domain that is generated in

depth. The second was the forward-modeling domain that was built in two-way

travel time. For this real data application, the vertical sampling of the rock physics

simulation grid (depth) was fixed to one meter and the two-ways travel time

sampling for the forward model grid was set to one millisecond. The depth and

time grids were selected to minimize the resampling effect on the well-log data.

Five million realizations were performed and 320 models with normalized

α≤0.88 were selected. The comparison between the real traces and the 320

solutions for 0° and 32° incident angles are shown in Figure 6.8.

85
Angle Traces at 0o
2

Amplitude [Dimensionless]
50 1
Time [msec]
100 0

150 -1

200 -2
50 100 150 200 250 300
Number of Traces
Angle Traces at 32o
2

Amplitude [Dimensionless]
50 1
Time [msec]

100 0

150 -1

200 -2
50 100 150 200 250 300
Number of Traces

Figure 6.8: Comparison between real angle traces (upper and lower left) and the
320 solutions for 0° (upper right) and 32° (lower right) incident angles.

Figure 6.9 shows that the similarity among the traces is large enough to

assume that all of 320 model traces were generated from probable geological

models. The upper and lower sands are well-delimited by most of the probable

inversion solutions. However, the gas-LSGS fluid contact can be located at any

depth because of the impossibility of elastically discriminating between gas and

low saturation gas sands. The possible models plot shows that all the possible

solutions for the LSGS-brine sand fluid contact have the fluid interface located in

the lower sand. In fact, most solutions have the fluid contact situated more than

20 meters below the top of the lower sand.

86
Figure 6.9: Probable models with α≤0.88. Grey, black and white represent
shales, gas sands and low saturation gas sands, respectively.

We are interested in the prediction of total porosity, shale volume, and

effective porosity prediction from the seismic inversion, as has been mentioned.

Figures 6.10, 6.11 and 6.12 show the total porosity, volume of shale, and

effective porosity for the 320 solutions. The elastic models shown in Figure 6.13

demonstrate that there are almost no differences in the elastic properties of the

gas sands and low saturation gas sands. In fact, the main differences are found

in the density rather than in the elastic moduli.

87
Figure 6.10: Total porosity of the 320 probable models.

Figure 6.11: Volume of shale of the 320 probable models.

88
Figure 6.12: Effective porosity of the 320 probable models.

Figure 6.13: Possible elastic models with α≤0.88.

89
6.5. Analysis

The combination of the probable models for the target zone was used for

the analysis of rock properties. Figure 6.14 shows the probability of the sand plot;

larger values mean higher possibility of having a sand sample at a given depth. It

can be observed that the upper sand has less probability than the lower sand,

due to its sand/shale interbedded geometry. We can assume any probability

cutoff to define the most probable sand/shale geometry; however, it is important

to mention that the larger the cutoff the smaller the sand fraction in the analyzed

interval; we chose a conservative value of 0.2 as the threshold for this real data

set. The comparison between the real and the estimated top values for

probability equal to or larger than 0.2 is shown in table 6.1. The comparison

confirms the accuracy of our proposed inversion scheme for lithology prediction.

90
20

40

60

80

100
Depth [m]

120

140

160

180

200

220
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
Probability of Sand

Figure 6.14: Probability of sand vertical plot.

Table 6.1: Comparison between the real and the estimated lithology top values
for probability larger than 0.2.

Inverted Tops [m] Real Tops [m]


Top Upper Sand 80 79
Base of Upper
103 98
Sand
Top Lower Sand 116 114
Base of Lower
176 181
Sand

91
Based on the inversion solutions for total porosity and effective porosity,

we generated the most probable vertical plot for both rock properties (Figure

6.15). The comparison shows that the inverted porosities underestimated the

larger porosity values and overestimated the lower observed well log porosity.

This smoothing effect is stronger in the upper sand where extreme property

values are further apart.

0 0

20 20

40 40

60 60

80 80

100 100
Depth [m]

Depth [m]

120 120

140 140

160 160

180 180

200 200

220 220
0 0.1 0.2 0.3 0.4 0.5 0 0.1 0.2 0.3 0.4 0.5
Total Porosity [fraction] Effective Porosity [fraction]

Figure 6.15: Total porosity and effective porosity comparison. The black lines
correspond to the well log information and the blue lines to the inverted rock
properties.

92
The most-likely fluid geometry (Figure 6.16) was generated from the 320

probable models section (Figure 6.9) and constrained by the most expected sand

profile (Figure 6.14). The comparison between the real fluid interfaces and the

inverted fluid contact depths (Table 6.2) shows a surprisingly close agreement

with the well data.

Figure 6.16: Most expected fluid seismic lithofacies (left) and probability of fluid
type (right).

Table 6.2: Comparison between the real and the estimated fluid contact depths.

Inverted Contacts [m] Real Contacts [m]


Gas - LSGS Contact 143 147
LSGS - Brine 168 176

93
Finally, the elastic properties probable models generated from the

stochastic inversion are compared with the Well-1 elastic logs in Figure 6.17. The

results of the inversion are remarkable similar to the well log measured data.

However, the S-wave velocity of the interbedded shales within the upper sand

section was not well predicted. These shales have a very low Vp/Vs ratio

compared to the rest of the shales.

0 0 0

20 20 20

40 40 40

60 60 60

80 80 80

100 100 100


Depth [m]

Depth [m]

Depth [m]
120 120 120

140 140 140

160 160 160

180 180 180

200 200 200

220 220 220


1.5 2 2.5 3 3.5 1.5 2 2.5 1 1.2 1.4 1.6 1.8
Vp [km/sec] Rhob [g/cc] Vs [km/sec]

Figure 6.17: Comparison between the most expected elastic models (blue line)
against well logs (black lines).

94
6.6. Conclusions

The outcomes of the 1D real data application of the proposed stochastic

inversion are remarkable similar to the available well log measured data. It is

important to emphasize that Well-1 was not used to constrain the inversion; it

was only used to populate the model space. The mean of the inverted rock

properties tends to be smoother than the measured values; in fact, the most

expected inverted properties have a tendency to underestimate larger rock

property values and overestimate lower property measurements. This

‘smoothing’ effect is stronger in layers where extreme property values are further

apart. This effect is observed in all the estimated rock properties.

The S-wave velocity of the shale member of the interbedded upper sand

was underestimated by the inversion; this suggests that an additional seismic

lithofacies might be included in the model space. Finally, the proposed inversion

approach demonstrates the viability of the integration seismic data, well log

information, and rock physics models in a practical workflow to estimate rock

properties for hydrocarbon exploration or development.

95
7. 3D REAL DATA APPLICATION

7.1. Introduction

In this chapter, we introduce the first 3D application of a real data set

using the proposed inversion algorithm. The real data seismic inversion is

performed using the same data set presented in the Chapter 6. The main

difference between the 1D and 3D stochastic inversion algorithms is the use of

geostatistics and well log data to constrain the outcomes.

In the first part of the chapter, we introduce some basic concepts of

geostatistics that are used to create lateral continuity to the stochastic seismic

inversion outcomes. The second part of the chapter shows the application of the

proposed 3D stochastic inversion algorithm to the 2D line of prestack seismic

data extracted from a 3D volume and supported by the Well-1 information.

Finally, we discuss the convergence criteria and the efficiency of the proposed

inversion algorithm.

7.2. The Kriging Estimator

Geostatistics is a set of mathematical tools used to describe the lateral

continuity of natural phenomena. Geostatistics was originally developed to

estimate probability distributions of mineral concentrations for mining

prospecting. Classical geostatistics is strongly related to the Kriging interpolation

algorithm (Isaaks and Srivastava, 1989). The Kriging algorithm is a family of

96
linear unbiased estimators that predict the value of a property in any given

location by computing a weighted average of the observed values of the studied

properties and their spatial distributions. All the Kriging algorithms assume that

the observed data are the result of a random process; then the objective of the

Kriging estimator is to minimize the error variance ( ).

The Kriging estimators are defined by the treatment of the trend

component (mean). In this particular application, we used the Ordinary Kriging to

include lateral continuity to the stochastic inversion. Ordinary Kriging assumes

that the local mean may vary greatly over the studied area. In this case, the

system of equations is:

∑ ( ) ( ) ( ) ( )
{ , (7.1)
∑ ( ) ( )

where:

uα, uβ = Location vectors for known location,

u = Location of the unknown location,

h = uα-uβ,

CR(h) = Covariance function,

λα(u) = Kriging weight.

In general, the weight of each data point decreases with increasing

distance to the location that is being estimated. The weights λα are calculated

from the covariance function that is generally derived from the semivariogram

97
model. The semivariogram model is a function that describes the degree of

spatial dependence of the target property. It is defined as the variance of the

difference between known values at two locations across the target area. The

Kriging estimator needs valid semivariogram models to compute the covariance

at any distance h. In isotropic environments, the correlation between the spatial

process at two point pairs p1 and p2 only depends on the separation distance h

and not on the orientation of the two points. However, real data often show

directional effects. Particularly in geologic applications, measurements along a

particular direction may be highly correlated, while the perpendicular direction

shows little or no correlation. This phenomenon is known as anisotropy.

There are two types of anisotropy. The simplest type occurs when the

covariance form is the same in all directions but the range S changes. In this

case, there is a single sill, but the semivariogram reaches the sill in a shorter lag

distance along a certain direction. In anisotropic settings h can be estimated

using the following relation:

( ), (7.2)

where R is the ratio of between the larger r1and the lower seal r2:

. (7.3)

In this real data application we use a spherical semivariogram to estimate

the Ordinary Kriging weights. The spherical semivariograms are defined as:

98
( ) ( ) ( ) , (7.4)

where h is the distance vector and a is a fitting constant defined from the

observed data.

7.3. 3D Real Data Application

As has been mentioned before, the 3D application integrates the 1-D

stochastic inversion algorithm with a Kriging engine to produce the lateral

constraint rock properties model. Figure 7.1 shows the simplified stochastic

inversion algorithm; the inversion scheme includes an additional choice to select

between the Markov Chain simulation for rock properties or the geostatistical

estimation of rock properties. In this real test we inverted the 50 traces that

constitute the 2D seismic line (Figure 6.2) using the Well-1 as initial well data and

the statistical rock physics model shown. To estimate the covariance we used an

anisotropic spherical semivariogram with a 0.1 nugget, a sill of 0.9 and ranges of

500 meters and 50 meters in horizontal and vertical directions respectively

(Figure 7.2). We chose the horizontal range as the maximum distance to accept

a trace for the geostastistical estimation of the rock properties.

99
Figure 7.1: Simplified 3D stochastic inversion workflow.

Figure 7.2: Anisotropic semivariogram model with a 0.1 nugget, a sill of 0.9 and
ranges of 500 meters and 50 meters.

100
Figure 7.3 shows the evolution of the variance model for one of the

stochastic simulation outcomes. The order selection of the seismic traces plays

an important role in the stochastic inversion process. To assure a random

selection of the trace’s order we implemented the Fisher and Yates shuffle

algorithm supported by a Markov Chain for each stochastic realization.

Therefore, any trace selection path should be independent from the previous

realizations. The upper section of Figure 7.3 shows the initial variance model, it is

clear that the closer to the Well-1 the smaller the variance. The lower variance

values are situated at the well location where the variance becomes zero. The

second, third and fourth sections (from top to bottom) represent three more

iterations that produced three new wells. Every new iteration reduced the

variance model (reducing the uncertainty) until the variance model was zero at

the fiftieth iteration.

The inversion was performed for the four seismic lithofacies defined in

Chapter 6: shales, gas sands, LSGS and brine sands. Figure 7.4 shows three

possible models generated by the proposed inversion algorithm. The total

porosity, volume of shale, P-wave velocity, bulk density and S-wave velocity

inversion outcomes are shown in figures 7.5 to 7.9, respectively. All the shown

models are probable and similar. However, they have small variations in the

depth of the seismic lithofacies interfaces, especially, in the gas sands and LSGS

interfaces where the impedance contrast is very small.

101
Figure 7.3: Variance model evolution for one of the stochastic simulation
outcomes.

Figure 7.4: Three possible stochastic models. Shales are represented by facies
1, gas sands by facies 2, LSGS by facies 3 and brine sands by facies 4.

102
Figure 7.5: Three possible stochastic total porosity models.

Figure 7.6: Three possible stochastic volume of shale models.

103
Figure 7.7: Three possible stochastic P-wave impedance models.

Figure 7.8: Three possible stochastic bulk density models.

104
Figure 7.9: Three possible stochastic S-wave velocity models.

A set of four hundred possible outcomes was generated from the seismic

section. This set of seismic inversions was used to calculate the most expecting

values of the estimated rock properties for each sample of the 2D seismic

section. As it is expected, these inverted rock property sections were smoother

versions of the original inverted outcomes.

Additionally, we generated the variance section for each rock property

inverted vertical section. The variance sections were used to better understand

the uncertainty associated with the inversion outcomes. Figure 7.10 shows the

most expected section of total porosity (upper section) and its associated

variance (lower section). Well-1 measured total porosity log is located in the trace

number 37. It can be observed that there is good agreement between the

measured data and the inverted section. However, Well-1 has a larger vertical

105
resolution than the inverted traces. As we observed in Chapter 6, the median of

the stochastic inversion outcomes tends to slightly underestimate larger porosity

values and overestimate lower porosity values. The variance section shows that

the upper gas sand has larger estimation uncertainty. This is due to the

interbedded gas-shale geometry of the upper sand. The zero variance trace

corresponds to the Well-1 location. The same phenomenon is observed in all the

rock property vertical sections; the upper sand has larger estimation errors than

the lower sand (Figures 7.11 to 7.14). In the bulk density inverted section (Figure

7.13) the variance of the upper sand and the lower sand gas intervals has larger

estimation errors than those of the LSGS and brine sand. This is caused by the

low predictability of the gas sands and LSGS fluid contact.

Figure 7.10: Most expected total porosity model generated from 400 stochastic
models.

106
Figure 7.11: Most expected volume of shale model generated from 400
stochastic models.

Figure 7.12: Most expected volume of P-wave velocity generated from 400
stochastic models.

107
Figure 7.13: Most expected volume of density generated from 400 stochastic
models.

Figure 7.14: Most expected volume of S-wave velocity generated from 400
stochastic models.

108
7.4. Convergence Criteria

A critical parameter for the proposed stochastic inversion execution is the

misfit threshold α. For this real data application we chose the L2-norm value

obtained from the synthetic tie as value of α. This selection was based on the

idea that the forward model operator is imperfect and cannot perfectly reproduce

real seismic data. Other sources of misfits, besides the forward modeling

operator, come from imperfections in the seismic data acquisition/processing and

deficiencies in the time-to-depth conversion model. It is important to point out that

the smaller α becomes, the more realizations are needed to obtain a statistically

significant number of probable models. Figure 7.15 shows 150,000 models

randomly selected from the five million realizations performed for the real

example inversion. It can be observed that only a few hundreds models look like

the studied real data set inverted.

Figure 7.15: Models randomly selected from the real dataset inversion.

109
The synthetic traces generated from the models shown in Figure 7.15 are

shown in Figure 7.16. For other shown models, the calculated seismic traces are

very noisy, without recognizable seismic interfaces. For the 150,000 traces

shown in Figure 7.16, only nine were selected as probable solutions. The

selection of α value smaller than the similarity obtained from the well log tie

would have deteriorated the inversion outcome, because the possible solutions

would have been generated from a partially sampled model space.

Figure 7.16: Comparison between real angle trace (right) and 150,000 solutions
(left) randomly selected from the five million realizations performed for the real
example inversion.

110
7.5. Algorithm Efficiency

The proposed algorithm was extensively tested in a Windows based

personal computer with two processors and eight cores. The results showed that

the amount of resources (computing time) needed by the our proposed algorithm

increase linearly (O(n)) with the number of tested models. The average running

time is 269 models/sec.

Figure 7.17: Algorithm efficiency, the proposed algorithm increase linearly with
the number of tested models.

111
6.7. Conclusions

The applicability of our proposed 3D stochastic inversion algorithm to real

case studies was demonstrated. The outcome of the inversion was a set of key

rock property sections that can be directly used for the evaluation of the

hydrocarbon potential of a given the inverted subsurface interval. Moreover, the

inversion solutions were generated directly in the depth domain. Therefore, with

our solutions, no time-to-depth is needed during the inversion process.

Results of the 3D inversion showed good vertical resolution and adequate

lateral continuity. As we observed in the 1D real data application, the real data

test of the 3D stochastic inversion shows a remarkable similarity to the available

well log data. However, the median of the inverted traces has lower vertical

variability than the Well-1 logs. Larger estimation errors were found in the upper

gas sand, due to its interbedded geometry.

The real-data tests showed that the similarity factor α is the most sensible

parameter in the proposed inversion algorithm. The selection of α value smaller

than the similarity obtained from the well log tie will deteriorate the inversion

outcome, since the possible solutions will be generated from a partially sampled

model space.

112
REFERENCES

Aki, K., and Richards, P. G., 1980, Quantitative seismology: theory and methods:
W. N. Treeman & Co. pp. 119-149.

Arpat, B. G., 2005, Sequential simulation with patterns: Ph.D. dissertation,


Stanford Univ. pp. 7-11.

Avseth, P., 2000, Combining rock physics and sedimentology for seismic
reservoir characterization in North Sea turbidite systems: Ph.D. dissertation,
Stanford Univ. pp. 14-49.

Avseth, P., Mukerji, T., and Mavko, G., 2005, Quantitative seismic interpretation:
applying rock physics tools to reduce interpretation risk: Cambridge University
Press. pp. 42-47.

Backus, G. E., 1962, Long-wave elastic anisotropy produced by horizontal


layering: J. Geophys. Res., 67, 4427-4440.

Balch, A. H., 1971, Color sonagrams - A new dimension in seismic data


interpretation: Geophysics, 36, 1074-1098.

Banchs, R. and Michelena, R., 2000, Well log estimates and confidence intervals
by using artificial neural networks: 70th Ann. Internat. Mtg,. Soc. of Expl.
Geophys., Expanded Abstracts, 1430-1432.

Batzle, M., and Wang., Z., 1992, Seismic properties of pore fluids: Geophysics,
57, 1396-1408.

Berryman, J. G., Berge, P. A., Bonner, B. P., 2002, Estimating rock porosity and
fluid saturation using only seismic velocities: Geophysics, 67, 391-404.

Bortoli, L. J., Alabert, F. A., Haas, A., and Journel, A. G., 1993, Constraining
Stochastic Images to Seismic Data, in A. Soares ed., Geostatistics Tróia 1992,
proc. 4th Inter. Geostat. Congr., Kluwer Academic Publ., 325-337.

Bosch, M., 1999, Lithologic tomography: From plural geophysical data to


lithology estimation: Journal of Geophysical Research, 104, 749–766.

Castagna, J. P., Batzle, M. L., and Eastwood, R. L., 1985, Relationships between
compressional-wave and shear-wave velocities in clastic silicate rocks:
Geophysics, 50, 571-581.

113
Castagna, J. P., Batzle, M. L., and Kan, T. K., 1993, Rock physics – The link
between rock properties and AVO response, in J. P. Castagna and M. Backus,
eds., Offset-dependent reflectivity – Theory and practice of AVO analysis: Soc. of
Expl. Geophys., 135-171.

Connolly, P., 1998, Calibration and inversion of non-zero offset seismic: 68th
Ann. Internat. Mtg,. Soc. of Expl. Geophys., Expanded Abstracts, 182-184.

Connolly, P., 1999, Elastic impedance: The Leading Edge, 18, 438-452.

Deutsch C., and Journel, A., 1998, GSLIB: Geostatistical Software Library:
Oxford University Press, 2nd edition.

Domenico, S. N., 1976, Effect of brine-gas mixture on velocity in an


unconsolidated sand reservoir: Geophysics, 41, 882-894.

Dvorkin, J., and Nur, A., 1996, Elasticity of High-Porosity Sandstones: Theory for
Two North Sea Datasets, Geophysics, 61, 1363-1370.

Dvorkin, J., Nur, A., and Yin, H., 1994, Effective Properties of Cemented
Granular Materials, Mechanics of Materials, 18, 351-366.

Fisher, A., Yates, F., 1948, Statistical tables for biological, agricultural and
medical research: Oliver & Boyd. Pp. 26-27.

Florez, J., 2005, Integrating geology, rock physics, and seismology for reservoir
quality prediction: Ph.D. dissertation, Stanford Univ. pp. 13-44.

Gardner, G., Gardner, L., and Gregory, A., 1974, Formation velocity and density
– The diagnostic basic for stratigraphic traps: Geophysics, 39, 770-780.

Gassmann, F., 1951, Über die elastizität poröser medien: Veirteljahrsschrift der
Naturforschenden Gesellschaft in Zürich, 96, 1-23.

Gonzalez, E. F., Mukerji, T. and Mavko, G., 2000, Facies identification using P-
to-P and P-to-S AVO attributes: 70th Ann. Internat. Mtg,. Soc. of Expl. Geophys.,
Expanded Abstracts, 98-101.

Gonzalez, E., 2006, Physical and quantitative Interpretation of seismic attributes


for rock and fluid interpretation: Ph.D. dissertation, Stanford Univ. pp. 49-78.

Goodway, B., Chen, T. and Downton, J., 1997, Improved AVO fluid detection and
lithology discrimination using Lame petrophysical parameters; “λρ”, “μρ”, & “λ/μ

114
fluid stack”, from P and S inversions: 67th Ann. Internat. Mtg,. Soc. of Expl.
Geophys., Expanded Abstracts, 183-186.

Grechka, V., 2008, Fluid substitution in porous and fractured solids: The non-
interaction approximation and Gassmann theory: International Journal of
Fracture, 148, 103-107.

Gray, D., 2002, Elastic inversion for Lamé parameters: 72th Ann. Internat. Mtg,.
Soc. of Expl. Geophys., Expanded Abstracts, 213-216.

Greenberg, M. L., and Castagna, J.P., 1992, Shear-wave velocity estimation in


porous rocks: Theoretical formulation, preliminary verification and applications,
Geophys. Prospect., 40, 195-209.

Guardiano, F., and Srivastava, R., 1993, Multivariate geostatistics: Beyond


bivariate moments, in A. Soares ed., Geostatistics Tróia 1992, proc. 4th Inter.
Geostat.Congr., Kluwer Academic Publ., 133-144.

Gutierrez, M., 2001, Rock physics and 3-D seismic characterization of reservoir
heterogeneities to improve recovery efficiency: Ph.D. dissertation, Stanford Univ.

Haas, A. and Dubrule, O., 1994, Geostatistical inversion - A sequential method of


stochastic reservoir modeling constrained by seismic data: First Break, 12, 561-
569.

Han, D., 1986, Effects of porosity and clay content on acoustic properties of
sandstones and unconsolidated sediments: Ph.D. thesis, Stanford University.

Harris, D. A., Lewis, J. J. M., and Wallace, D. J., 1993, The identification of
lithofacies types in geological imagery using neural networks, Conf. papers vol.,
EUROCAIPEP 93, Aberdeen.

Hashin, Z., and Shtrikman, S., 1963, A variational approach to the elastic
behavior of multiphase materials: J. Mech. Phys. Solids, 11, 127-140.

Hastings, W.K., 1970, Monte Carlo sampling methods using Markov Chains and
their applications: Biometrika, 57, 97–109.

Hearst, J. R., Nelson, P. H., Pailet, F. L., 2000, Well logging for physical
properties: McGraw-Hill, 2nd edition.

115
Hirsche, K., Boerner, S., Kalkomey, C. and Gastaldi, C., 1998, Avoiding pitfalls in
geostatistical reservoir characterization: A survival guide: The Leading Edge, 17,
493-504.

Isaaks, E., and Srivastava, R., 1989, An introduction to applied geostatistics:


Oxford University Press. pp. 278-322.

Kalkomey, C. T., 1997, Potential risks when using seismic attributes as


predictors of reservoir properties: The Leading Edge, 16, 247-251.

Landrø, M., 1999, Discrimination between pressure and fluid saturation changes
from time lapse seismic data, 69th Ann. Int. Meet., Soc. Expl. Geophys.,
Expanded Abstracts, 1651-1654.

Mallick, S., 2001, Hybrid inversion, elastic impedance inversion, and prestack
waveform inversion; 71st Ann. Internat. Mtg: Soc. of Expl. Geophys., 706-709.
Marion, D., 1990, Acoustical, mechanical and transport properties of sediments
and granular materials: Ph.D. thesis, Stanford University.

Matteucci, G., 1996, Seismic attribute analysis and calibration: A general


procedure and a case study: 66th Ann. Internat. Mtg,. Soc. of Expl. Geophys.,
Expanded Abstracts, 373-376.

Mavko, G., and Jizba, D., 1991, Estimation grain-scale fluid effects on velocity
dispersion in rocks: Geophysics, 56, 1940-1949.

Mavko, G., Mukerji, T., and Dvorkin, J., 1998, The rock physics handbook:
Cambridge University Press.

Metropolis, N., Rosenbluth, A.W., Rosenbluth, M.N., Teller, A.H., Teller, E., 1953,
Equations of state calculations by fast computing machines: Journal of Chemical
Physics, 21, 6, 1087–1092.

Mukerji, T., Jørstad, A., Mavko, G., and Granli, J., 1998, Near and far offset
impedances: Seismic attributes for identifying lithofacies and pore fluids:
Geophysical Research Letter, 25, 4557-4560.

Mukerji, T., Avseth, P., Mavko, G, Takahashi, I., and Gonzalez, E. F., 2001,
Statistical rock physics: Combining rock physics, information theory, and
geostatistics to reduce uncertainty in seismic reservoir characterization: The
Leading Edge, 20, 313-319.

116
Oldenburg, D. W., Scheuer, T., and Levy, S., 1983, Recovery of the acoustic
impedance from reflection seismograms: Geophysics, 48, 1318-1337.

Ostrander, W. J., 1984, Plane-wave reflection coefficients for gas sands at non-
normal angles of incidence: Geophysics, 49, 1637-1648.

Papoulis, A., 1984, Probability, random variables, and stochastic processes:


McGraw-Hill.

Possato, S., Saito, M., Curtis, M. P., and Martinez, R. D., 1983, Interpretation of
three-dimensional seismic attributes contributes to stratigraphic analysis of
Pampo oil field: 53rd Ann. Internat. Mtg,. Soc. of Expl. Geophys., Expanded
Abstracts, session: S16.2.

Raymer, L. L., E. R. Hunt, and J. S. Gardner, 1980, An improved sonic transit


time-to-porosity transform: 21st Annual Logging Symposium Transaction, Society
of Professional Well Log Analysts, Paper P.

Ruiz, F., 2009, Porous grain model and equivalent elastic medium approach for
predicting effective elastic properties of sedimentary rocks: Ph.D. dissertation,
Stanford Univ.

Russell, B. H., 1988, Introduction to seismic inversion methods: Soc. of Expl.


Geophys. Course notes from SEG Continuing Education course.

Sengupta, M., 2000, Integrating rock physics and flow simulation to reduce
uncertainties in seismic reservoir monitoring: Ph.D. dissertation, Stanford Univ.

Shuey, R. T., 1985, A simplification of the Zoeppritz equations: Geophysics, 50,


609-614.

Sheriff, R., 1989, Geophysical Methods: Prentice Hall.

Takahashi, I., 2000, Quantifying information and uncertainty of rock property


estimation from seismic data: Ph.D. dissertation, Stanford Univ.

Taner, M. T., Koehler, F., and Sheriff, R. E., 1979, Complex seismic trace
analysis: Geophysics, 44, 1041-1063.

Tarantola, A., 2005, Inverse problem theory and methods for model parameter
estimation: Society for Industrial and Applied Mathematics.

Thomsen, L., 1985, Biot-consistent elastic moduli of porous rocks: Low-


frequency limit: Geophysics, 50, 2797-2807.

117
Tonn, R., 1992, Statistical approach to correlate reservoir parameters and 3-D
seismic attributes, 62nd Ann. Internat. Mtg,. Soc. of Expl. Geophys., Expanded
Abstracts, 272-274.

Wu, Y., 2000, Estimation of gas saturation using P-to-S converted waves; 70th
Ann. Internat. Mtg,. Soc. of Expl. Geophys., Expanded Abstracts, 158-161.

Wyllie, M. R., Gregory, A. R., and Gardner, L. W., 1956: Elastic wave velocities in
heterogeneous and porous media, Geophysics, 21, 41-70.
Yin, H., Mavko, G., and Nur, A., 1993, Critical porosity; a physical boundary in
poroelasticity: 34th U. S. symposium on Rock mechanics, in Haimson, B. C., ed.,
Rock mechanics in the 1990s, Oxford-New York, International Journal of Rock
Mechanics and Mining Sciences & Geomechanics, Abstracts, 30, 805-808.

Zeng, H., Backus, M. M., Barrow, K. T. and Tyler, N., 1996, Facies mapping from
three-dimensional seismic data: Potential and guidelines from a Tertiary
sandstone-shale sequence model, Powderhorn Field, Calhoun County, Texas:
AAPG Bull., 80, 16-46.

118

Vous aimerez peut-être aussi